You are on page 1of 74

a

r
X
i
v
:
1
1
0
4
.
1
1
0
6
v
2


[
m
a
t
h
.
D
G
]


7

A
p
r

2
0
1
1
Lecture Notes in Lie Groups
Vladimir G. Ivancevic

Tijana T. Ivancevic

Abstract
These lecture notes in Lie Groups are designed for a 1semester third year
or graduate course in mathematics, physics, engineering, chemistry or biology.
This landmark theory of the 20th Century mathematics and physics gives a
rigorous foundation to modern dynamics, as well as eld and gauge theories
in physics, engineering and biomechanics. We give both physical and medical
examples of Lie groups. The only necessary background for comprehensive
reading of these notes are advanced calculus and linear algebra.
Contents
1 Preliminaries: Sets, Maps and Diagrams 3
1.1 Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Commutative Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . 5
2 Groups 6
3 Manifolds 8
3.1 Denition of a Manifold . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2 Formal Denition of a Smooth Manifold . . . . . . . . . . . . . . . . 12
3.3 Smooth Maps Between Smooth Manifolds . . . . . . . . . . . . . . . 13
3.4 Tangent Bundle and Lagrangian Dynamics . . . . . . . . . . . . . . 14
3.5 Cotangent Bundle and Hamiltonian Dynamics . . . . . . . . . . . . . 17

Land Operations Division, Defence Science & Technology Organisation, P.O. Box 1500, Edin-
burgh SA 5111, Australia (e-mail: Vladimir.Ivancevic@dsto.defence.gov.au)

Tesla Science Evolution Institute & QLIWW IP Pty Ltd., Adelaide, Australia (e-mail:
tijana.ivancevic@alumni.adelaide.edu.au)
1
4 Lie Groups 18
4.1 Denition of a Lie Group . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Lie Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3 One-Parameter Subgroup . . . . . . . . . . . . . . . . . . . . . . . . 24
4.4 Exponential Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.5 Adjoint Representation . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.6 Actions of Lie Groups on Smooth Manifolds . . . . . . . . . . . . . . 26
4.7 Basic Tables of Lie Groups and Their Lie Algebras . . . . . . . . . . 28
4.8 Representations of Lie groups . . . . . . . . . . . . . . . . . . . . . . 31
4.9 Root Systems and Dynkin Diagrams . . . . . . . . . . . . . . . . . . 32
4.9.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.9.2 Classication . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.9.3 Dynkin Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.9.4 Irreducible Root Systems . . . . . . . . . . . . . . . . . . . . 36
4.10 Simple and Semisimple Lie Groups and Algebras . . . . . . . . . . . 37
5 Some Classical Examples of Lie Groups 38
5.1 Galilei Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.2 General Linear Group . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.3 Rotational Lie Groups in Human/Humanoid Biomechanics . . . . . 40
5.3.1 Uniaxial Group of Joint Rotations . . . . . . . . . . . . . . . 41
5.3.2 ThreeAxial Group of Joint Rotations . . . . . . . . . . . . . 43
5.3.3 The Heavy Top . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.4 Euclidean Groups of Rigid Body Motion . . . . . . . . . . . . . . . . 45
5.4.1 Special Euclidean Group SE(2) in the Plane . . . . . . . . . 46
5.4.2 Special Euclidean Group SE(3) in the 3D Space . . . . . . . 47
5.5 Basic Mechanical Examples . . . . . . . . . . . . . . . . . . . . . . . 50
5.5.1 SE(2)Hovercraft . . . . . . . . . . . . . . . . . . . . . . . . 50
5.5.2 SO(3)Satellite . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.5.3 SE(3)Submarine . . . . . . . . . . . . . . . . . . . . . . . . 52
5.6 NewtonEuler SE(3)Dynamics . . . . . . . . . . . . . . . . . . . . 52
5.6.1 SO(3) : Euler Equations of Rigid Rotations . . . . . . . . . . 52
5.6.2 SE(3) : Coupled NewtonEuler Equations . . . . . . . . . . . 53
5.7 Symplectic Group in Hamiltonian Mechanics . . . . . . . . . . . . . 56
6 Medical Applications: Prediction of Injuries 57
6.1 General Theory of MusculoSkeletal Injury Mechanics . . . . . . . . 57
6.2 Analytical Mechanics of Traumatic Brain Injury (TBI) . . . . . . . . 63
6.2.1 The SE(3)jolt: the cause of TBI . . . . . . . . . . . . . . . 63
6.2.2 SE(3)group of brains micromotions within the CSF . . . 64
2
6.2.3 Brains natural SE(3)dynamics . . . . . . . . . . . . . . . . 65
6.2.4 Brains traumatic dynamics: the SE(3)jolt . . . . . . . . . 68
6.2.5 Brains dislocations and disclinations caused by the SE(3)jolt 69
1 Preliminaries: Sets, Maps and Diagrams
1.1 Sets
Given a map (or, a function) f : A B, the set A is called the domain of f, and
denoted Domf. The set B is called the codomain of f, and denoted Codf. The
codomain is not to be confused with the range of f(A), which is in general only a
subset of B (see [8, 9]).
A map f : X Y is called injective, or 11, or an injection, i for every y in the
codomain Y there is at most one x in the domain X with f(x) = y. Put another way,
given x and x

in X, if f(x) = f(x

), then it follows that x = x

. A map f : X Y
is called surjective, or onto, or a surjection, i for every y in the codomain Codf
there is at least one x in the domain X with f(x) = y. Put another way, the range
f(X) is equal to the codomain Y . A map is bijective i it is both injective and
surjective. Injective functions are called monomorphisms, and surjective functions
are called epimorphisms in the category of sets (see below). Bijective functions are
called isomorphisms.
A relation is any subset of a Cartesian product (see below). By denition, an
equivalence relation on a set X is a relation which is reexive, symmetrical and
transitive, i.e., relation that satises the following three conditions:
1. Reexivity: each element x X is equivalent to itself, i.e., xx;
2. Symmetry: for any two elements a, b X, ab implies ba; and
3. Transitivity: ab and bc implies ac.
Similarly, a relation denes a partial order on a set S if it has the following
properties:
1. Reexivity: a a for all a S;
2. Antisymmetry: a b and b a implies a = b; and
3. Transitivity: a b and b c implies a c.
A partially ordered set (or poset) is a set taken together with a partial order on
it. Formally, a partially ordered set is dened as an ordered pair P = (X, ), where
X is called the ground set of P and is the partial order of P.
3
1.2 Maps
Let f and g be maps with domains A and B. Then the maps f +g, f g, fg, and
f/g are dened as follows (see [8, 9]):
(f +g)(x) = f(x) +g(x) domain = A B,
(f g)(x) = f(x) g(x) domain = A B,
(fg)(x) = f(x) g(x) domain = A B,
_
f
g
_
(x) =
f(x)
g(x)
domain = x A B : g(x) ,= 0.
Given two maps f and g, the composite map f g, called the composition of f
and g, is dened by
(f g)(x) = f(g(x)).
The (f g)machine is composed of the gmachine (rst) and then the fmachine,
x [[g]] g(x) [[f]] f(g(x)).
For example, suppose that y = f(u) =

u and u = g(x) = x
2
+ 1. Since y is a
function of u and u is a function of x, it follows that y is ultimately a function of x.
We calculate this by substitution
y = f(u) = f g = f(g(x)) = f(x
2
+ 1) =
_
x
2
+ 1.
If f and g are both dierentiable (or smooth, i.e., C

) maps and h = f g is the


composite map dened by h(x) = f(g(x)), then h is dierentiable and h

is given by
the product:
h

(x) = f

(g(x)) g

(x).
In Leibniz notation, if y = f(u) and u = g(x) are both dierentiable maps, then
dy
dx
=
dy
du
du
dx
.
The reason for the name chain rule becomes clear if we add another link to the chain.
Suppose that we have one more dierentiable map x = h(t). Then, to calculate the
derivative of y with respect to t, we use the chain rule twice,
dy
dt
=
dy
du
du
dx
dx
dt
.
Given a 11 continuous (i.e., C
0
) map F with a nonzero Jacobian

(x,...)
(u,...)

that
maps a region S onto a region R, we have the following substitution formulas:
4
1. For a single integral,
_
R
f(x) dx =
_
S
f(x(u))
x
u
du;
2. For a double integral,
__
R
f(x, y) dA =
__
S
f(x(u, v), y(u, v))

(x, y)
(u, v)

dudv;
3. For a triple integral,
___
R
f(x, y, z) dV =
___
S
f(x(u, v, w), y(u, v, w), z(u, v, w))

(x, y, z)
(u, v, w)

dudvdw;
4. Generalization to ntuple integrals is obvious.
1.3 Commutative Diagrams
Many properties of mathematical systems can be unied and simplied by a presen-
tation with commutative diagrams of arrows. Each arrow f : X Y represents a
function (i.e., a map, transformation, operator); that is, a source (domain) set X, a
target (codomain) set Y , and a rule x f(x) which assigns to each element x X
an element f(x) Y . A typical diagram of sets and functions is (see [8, 9]):
X Y

f
h

g or
X f(X)

f
h

g(f(X))

g
This diagram is commutative i h = gf, where gf is the usual composite function
g f : X Z, dened by x g(f(x)).
Less formally, composing maps is like following directed paths from one object to
another (e.g., from set to set). In general, a diagram is commutative i any two paths
along arrows that start at the same point and nish at the same point yield the same
homomorphism via compositions along successive arrows. Commutativity of the
whole diagram follows from commutativity of its triangular components. Study of
commutative diagrams is popularly called diagram chasing, and provides a powerful
tool for mathematical thought.
Many properties of mathematical constructions may be represented by universal
properties of diagrams. Consider the Cartesian product XY of two sets, consisting
as usual of all ordered pairs x, y of elements x X and y Y . The projections
5
x, y x, x, y y of the product on its axes X and Y are functions p :
X Y X, q : X Y Y . Any function h : W X Y from a third set W
is uniquely determined by its composites p h and q h. Conversely, given W and
two functions f and g as in the diagram below, there is a unique function h which
makes the following diagram commute:
X X Y

p
Y

q
W
f


h
g

This property describes the Cartesian product X Y uniquely; the same diagram,
read in the category of topological spaces or of groups, describes uniquely the Carte-
sian product of spaces or of the direct product of groups.
2 Groups
A group is a pointed set (G, e) with a multiplication : GG G and an inverse
: G G such that the following diagrams commute (see [10, 8, 9]):
1.
G
1

G GG

(e, 1)
G

(1, e)

(e is a twosided identity)
2.
GG G

GGG GG

(associativity)
6
3.
G
e

G GG

(, 1)
G

(1, )

(inverse).
Here e : G G is the constant map e(g) = e for all g G. (e, 1) means the map
such that (e, 1)(g) = (e, g), etc. A group G is called commutative or Abelian group
if in addition the following diagram commutes
GG GG

T
G

where T : G G G G is the switch map T(g


1
, g
2
) = (g
2
, g
1
), for all (g
1
, g
2
)
GG.
A group G acts (on the left) on a set A if there is a function : G A A
such that the following diagrams commute:
1.
A GA

(e, 1)
1

2.
GA A

GGA GA

where (e, 1)(x) = (e, x) for all x A. The orbits of the action are the sets
Gx = gx : g G for all x A.
7
Given two groups (G, ) and (H, ), a group homomorphism from (G, ) to (H, )
is a function h : G H such that for all x and y in G it holds that
h(x y) = h(x) h(y).
From this property, one can deduce that h maps the identity element e
G
of G to
the identity element e
H
of H, and it also maps inverses to inverses in the sense that
h(x
1
) = h(x)
1
. Hence one can say that h is compatible with the group structure.
The kernel Ker h of a group homomorphism h : G H consists of all those
elements of G which are sent by h to the identity element e
H
of H, i.e.,
Ker h = x G : h(x) = e
H
.
The image Imh of a group homomorphism h : G H consists of all elements
of G which are sent by h to H, i.e.,
Imh = h(x) : x G.
The kernel is a normal subgroup of G and the image is a subgroup of H. The
homomorphism h is injective (and called a group monomorphism) i Ker h = e
G
,
i.e., i the kernel of h consists of the identity element of G only.
3 Manifolds
A manifold is an abstract mathematical space, which locally (i.e., in a closeup
view) resembles the spaces described by Euclidean geometry, but which globally
(i.e., when viewed as a whole) may have a more complicated structure (see [11]).
For example, the surface of Earth is a manifold; locally it seems to be at, but viewed
as a whole from the outer space (globally) it is actually round. A manifold can be
constructed by gluing separate Euclidean spaces together; for example, a world
map can be made by gluing many maps of local regions together, and accounting
for the resulting distortions.
As main puremathematical references for manifolds we recommend popular
graduate textbooks by two exBourbaki members, Serge Lang [13, 12] and Jean
Dieudonne [14, 15]. Besides, the reader might wish to consult some other classics,
including [11, 16, 17, 18, 19, 20, 3]. Finally, as rstorder applications, we recom-
mend three popular textbooks in mechanics, [2, 1, 4], as well as our own geometrical
monographs [8, 9].
Another example of a manifold is a circle S
1
. A small piece of a circle appears to
be like a slightlybent part of a straight line segment, but overall the circle and the
8
segment are dierent 1D manifolds. A circle can be formed by bending a straight
line segment and gluing the ends together.
1
The surfaces of a sphere
2
and a torus
3
are examples of 2D manifolds. Manifolds
1
Locally, the circle looks like a line. It is 1D, that is, only one coordinate is needed to say
where a point is on the circle locally. Consider, for instance, the top part of the circle, where the
ycoordinate is positive. Any point in this part can be described by the xcoordinate. So, there is a
continuous bijection
top
(a mapping which is 11 both ways), which maps the top part of the circle
to the open interval (1, 1), by simply projecting onto the rst coordinate:
top
(x, y) = x. Such
a function is called a chart. Similarly, there are charts for the bottom, left , and right parts of the
circle. Together, these parts cover the whole circle and the four charts form an atlas (see the next
subsection) for the circle. The top and right charts overlap: their intersection lies in the quarter
of the circle where both the x and the ycoordinates are positive. The two charts
top
and

right
map this part bijectively to the interval (0, 1). Thus a function T from (0, 1) to itself can be
constructed, which rst inverts the top chart to reach the circle and then follows the right chart
back to the interval:
T(a) =
right
_

1
top
(a)
_
=
right
_
a,
_
1 a
2
_
=
_
1 a
2
.
Such a function is called a transition map. The top, bottom, left, and right charts show that the
circle is a manifold, but they do not form the only possible atlas. Charts need not be geometric
projections, and the number of charts is a matter of choice. T and the other transition functions
are dierentiable on the interval (0, 1). Therefore, with this atlas the circle is a dierentiable, or
smooth manifold.
2
The surface of the sphere S
2
can be treated in almost the same way as the circle S
1
. It can be
viewed as a subset of R
3
, dened by: S = |(x, y, z) R
3
[x
2
+y
2
+z
2
= 1. The sphere is 2D, so
each chart will map part of the sphere to an open subset of R
2
. Consider the northern hemisphere,
which is the part with positive z coordinate. The function dened by (x, y, z) = (x, y), maps
the northern hemisphere to the open unit disc by projecting it on the (x, y)plane. A similar chart
exists for the southern hemisphere. Together with two charts projecting on the (x, z)plane and two
charts projecting on the (y, z)plane, an atlas of six charts is obtained which covers the entire sphere.
This can be easily generalized to an nD sphere S
n
= |(x1, x2, ..., xn) R
n
[x
2
1
+x
2
2
+... +x
2
n
= 1.
An nsphere S
n
can be also constructed by gluing together two copies of R
n
. The transition
map between them is dened as R
n
\|0 R
n
\|0 : x x/|x|
2
. This function is its own inverse,
so it can be used in both directions. As the transition map is a (C

)smooth function, this atlas


denes a smooth manifold.
3
A torus (pl. tori), denoted by T
2
, is a doughnutshaped surface of revolution generated by
revolving a circle about an axis coplanar with the circle. The sphere S
2
is a special case of the
torus obtained when the axis of rotation is a diameter of the circle. If the axis of rotation does not
intersect the circle, the torus has a hole in the middle and resembles a ring doughnut, a hula hoop
or an inated tire. The other case, when the axis of rotation is a chord of the circle, produces a
sort of squashed sphere resembling a round cushion.
A torus can be dened parametrically by:
x(u, v) = (R +r cos v) cos u, y(u, v) = (R+r cos v) sin u, z(u, v) = r sin v,
where u, v [0, 2], R is the distance from the center of the tube to the center of the torus, and r
is the radius of the tube. According to a broader denition, the generator of a torus need not be a
circle but could also be an ellipse or any other conic section.
Topologically, a torus is a closed surface dened as product of two circles: T
2
= S
1
S
1
. The
9
are important objects in mathematics, physics and control theory, because they
allow more complicated structures to be expressed and understood in terms of the
wellunderstood properties of simpler Euclidean spaces (see [9]).
The Cartesian product of manifolds is also a manifold (note that not every man-
ifold can be written as a product). The dimension of the product manifold is the
sum of the dimensions of its factors. Its topology is the product topology, and a
Cartesian product of charts is a chart for the product manifold. Thus, an atlas
for the product manifold can be constructed using atlases for its factors. If these
atlases dene a dierential structure on the factors, the corresponding atlas denes
a dierential structure on the product manifold. The same is true for any other
structure dened on the factors. If one of the factors has a boundary, the product
manifold also has a boundary. Cartesian products may be used to construct tori
and cylinders, for example, as S
1
S
1
and S
1
[0, 1], respectively.
Manifolds need not be connected (all in one piece): a pair of separate circles
is also a topological manifold(see below). Manifolds need not be closed: a line
segment without its ends is a manifold. Manifolds need not be nite: a parabola is
a topological manifold.
Manifolds
4
can be viewed using either extrinsic or intrinsic view. In the extrinsic
view, usually used in geometry and topology of surfaces, an nD manifold M is seen
as embedded in an (n + 1)D Euclidean space R
n+1
. Such a manifold is called a
codimension 1 space. With this view it is easy to use intuition from Euclidean
spaces to dene additional structure. For example, in a Euclidean space it is always
clear whether a vector at some point is tangential or normal to some surface through
that point. On the other hand, the intrinsic view of an nD manifold M is an
abstract way of considering M as a topological space by itself, without any need
surface described above, given the relative topology from R
3
, is homeomorphic to a topological torus
as long as it does not intersect its own axis.
One can easily generalize the torus to arbitrary dimensions. An ntorus T
n
is dened as a
product of n circles: T
n
= S
1
S
1
S
1
. Equivalently, the ntorus is obtained from the
ncube (the R
n
generalization of the ordinary cube in R
3
) by gluing the opposite faces together.
An ntorus T
n
is an example of an nD compact manifold. It is also an important example of a
Lie group (see below).
4
Additional structures are often dened on manifolds. Examples of manifolds with additional
structure include:
dierentiable (or, smooth manifolds, on which one can do calculus;
Riemannian manifolds, on which distances and angles can be dened; they serve as the
conguration spaces in mechanics;
symplectic manifolds, which serve as the phase spaces in mechanics and physics;
4D pseudoRiemannian manifolds which model spacetime in general relativity.
10
for surrounding (n + 1)D Euclidean space. This view is more exible and thus it
is usually used in highdimensional mechanics and physics (where manifolds used
represent conguration and phase spaces of dynamical systems), can make it harder
to imagine what a tangent vector might be.
3.1 Denition of a Manifold
Consider a set M (see Figure 1) which is a candidate for a manifold. Any point
x M has its Euclidean chart, given by a 11 and onto map
i
: M R
n
, with its
Euclidean image V
i
=
i
(U
i
). More precisely, a chart
i
is dened by (see [8, 9])

i
: M U
i
x
i
(x) V
i
R
n
,
where U
i
M and V
i
R
n
are open sets.
Figure 1: Geometric picture of the manifold concept.
Clearly, any point x M can have several dierent charts (see Figure 1). Con-
sider a case of two charts,
i
,
j
: M R
n
, having in their images two open sets,
V
ij
=
i
(U
i
U
j
) and V
ji
=
j
(U
i
U
j
). Then we have transition functions
ij
between them,

ij
=
j

1
i
: V
ij
V
ji
, locally given by
ij
(x) =
j
(
1
i
(x)).
If transition functions
ij
exist, then we say that two charts,
i
and
j
are compat-
ible. Transition functions represent a general (nonlinear) transformations of coordi-
nates, which are the core of classical tensor calculus.
11
A set of compatible charts
i
: M R
n
, such that each point x M has its
Euclidean image in at least one chart, is called an atlas. Two atlases are equivalent
i all their charts are compatible (i.e., transition functions exist between them), so
their union is also an atlas. A manifold structure is a class of equivalent atlases.
Finally, as charts
i
: M R
n
were supposed to be 1-1 and onto maps, they
can be either homeomorphisms, in which case we have a topological (C
0
) manifold,
or dieomorphisms, in which case we have a smooth (C
k
) manifold.
3.2 Formal Denition of a Smooth Manifold
Given a chart (U, ), we call the set U a coordinate domain, or a coordinate neigh-
borhood of each of its points. If in addition (U) is an open ball in R
n
, then U
is called a coordinate ball. The map is called a (local ) coordinate map, and the
component functions (x
1
, ..., x
n
) of , dened by (m) = (x
1
(m), ..., x
n
(m)), are
called local coordinates on U [8, 9].
Two charts (U
1
,
1
) and (U
2
,
2
) such that U
1
U
2
,= are called compatible
if
1
(U
1
U
2
) and
2
(U
2
U
1
) are open subsets of R
n
. A family (U

)
A
of
compatible charts on M such that the U

form a covering of M is called an atlas.


The maps

(U

(U

) are called the transition maps,


for the atlas (U

)
A
, where U

= U

, so that we have a commutative


triangle:

(U

(U

An atlas (U

)
A
for a manifold M is said to be a C
k
atlas, if all transi-
tion maps

(U

(U

) are of class C
k
. Two C
k
atlases are called
C
k
equivalent, if their union is again a C
k
atlas for M. An equivalence class of
C
k
atlases is called a C
k
structure on M. In other words, a smooth structure on
M is a maximal smooth atlas on M, i.e., such an atlas that is not contained in any
strictly larger smooth atlas. By a C
k
manifold M, we mean a topological manifold
together with a C
k
structure and a chart on M will be a chart belonging to some
atlas of the C
k
structure. Smooth manifold means C

manifold, and the word


smooth is used synonymously with C

.
Sometimes the terms local coordinate system or parametrization are used
instead of charts. That M is not dened with any particular atlas, but with an
equivalence class of atlases, is a mathematical formulation of the general covariance
principle. Every suitable coordinate system is equally good. A Euclidean chart
12
may well suce for an open subset of R
n
, but this coordinate system is not to be
preferred to the others, which may require many charts (as with polar coordinates),
but are more convenient in other respects.
For example, the atlas of an nsphere S
n
has two charts. If N = (1, 0, ..., 0)
and S = (1, ..., 0, 0) are the north and south poles of S
n
respectively, then the two
charts are given by the stereographic projections from N and S:

1
: S
n
N R
n
,
1
(x
1
, ..., x
n+1
) = (x
2
/(1 x
1
), . . . , x
n+1
/(1 x
1
)), and

2
: S
n
S R
n
,
2
(x
1
, ..., x
n+1
) = (x
2
/(1 +x
1
), . . . , x
n+1
/(1 +x
1
)),
while the overlap map
2

1
1
: R
n
0 R
n
0 is given by the dieomorphism
(
2

1
1
)(z) = z/[[z[[
2
, for z in R
n
0, from R
n
0 to itself.
Various additional structures can be imposed on R
n
, and the corresponding man-
ifold M will inherit them through its covering by charts. For example, if a covering
by charts takes their values in a Banach space E, then E is called the model space
and M is referred to as a C
k
Banach manifold modelled on E. Similarly, if a cov-
ering by charts takes their values in a Hilbert space 1, then 1 is called the model
space and M is referred to as a C
k
Hilbert manifold modelled on 1. If not other-
wise specied, we will consider M to be an Euclidean manifold, with its covering by
charts taking their values in R
n
.
For a Hausdor C
k
manifold the following properties are equivalent: (i) it is
paracompact; (ii) it is metrizable; (iii) it admits a Riemannian metric;
5
(iv) each
connected component is separable.
3.3 Smooth Maps Between Smooth Manifolds
A map : M N between two manifolds M and N, with M m (m) N, is
called a smooth map, or C
k
map, if we have the following charting [8, 9]:
5
Recall the corresponding properties of a Euclidean metric d. For any three points x, y, z R
n
,
the following axioms are valid:
M1 : d(x, y) > 0, for x ,= y; and d(x, y) = 0, for x = y;
M2 : d(x, y) = d(y, x); M3 : d(x, y) d(x, z) +d(z, y).
13

U
m
V
(m)
M N

(U) (V )
(m) ((m))
R
m

R
n

This means that for each m M and each chart (V, ) on N with (m) V there
is a chart (U, ) on M with m U, (U) V , and =
1
is C
k
, that is,
the following diagram commutes:
(U) (V )

M U V N

Let M and N be smooth manifolds and let : M N be a smooth map. The


map is called a covering, or equivalently, M is said to cover N, if is surjective
and each point n N admits an open neighborhood V such that
1
(V ) is a union
of disjoint open sets, each dieomorphic via to V .
A C
k
map : M N is called a C
k
dieomorphism if is a bijection,

1
: N M exists and is also C
k
. Two manifolds are called dieomorphic if there
exists a dieomorphism between them. All smooth manifolds and smooth maps
between them form the category /.
3.4 Tangent Bundle and Lagrangian Dynamics
The tangent bundle of a smooth nmanifold is the place where tangent vectors live,
and is itself a smooth 2nmanifold. Vectorelds are cross-sections of the tangent
bundle. The Lagrangian is a natural energy function on the tangent bundle (see
[8, 9]).
14
In mechanics, to each nD conguration manifold M there is associated its 2nD
velocity phasespace manifold, denoted by TM and called the tangent bundle of
M (see Figure 2). The original smooth manifold M is called the base of TM.
There is an onto map : TM M, called the projection. Above each point
x M there is a tangent space T
x
M =
1
(x) to M at x, which is called a bre.
The bre T
x
M TM is the subset of TM, such that the total tangent bundle,
TM =
_
mM
T
x
M, is a disjoint union of tangent spaces T
x
M to M for all points
x M. From dynamical perspective, the most important quantity in the tangent
bundle concept is the smooth map v : M TM, which is an inverse to the projection
, i.e, v = Id
M
, (v(x)) = x. It is called the velocity vectoreld. Its graph
(x, v(x)) represents the crosssection of the tangent bundle TM. This explains
the dynamical term velocity phasespace, given to the tangent bundle TM of the
manifold M.
Figure 2: A sketch of a tangent bundle TM of a smooth manifold M.
If [a, b] is a closed interval, a C
0
map : [a, b] M is said to be dierentiable
at the endpoint a if there is a chart (U, ) at (a) such that the following limit exists
and is nite:
d
dt
( )(a) ( )

(a) = lim
ta
( )(t) ( )(a)
t a
. (1)
Generalizing (1), we get the notion of the curve on a manifold. For a smooth
manifold M and a point m M a curve at m is a C
0
map : I M from an
interval I R into M with 0 I and (0) = m.
Two curves
1
and
2
passing though a point m U are tangent at m with
respect to the chart (U, ) if (
1
)

(0) = (
2
)

(0). Thus, two curves are tangent


if they have identical tangent vectors (same direction and speed) in a local chart on
a manifold.
For a smooth manifold M and a point m M, the tangent space T
m
M to M at
m is the set of equivalence classes of curves at m:
T
m
M = []
m
: is a curve at a point m M.
15
A C
k
map : M m (m) N between two manifolds M and N induces
a linear map T
m
: T
m
M T
(m)
N for each point m M, called a tangent map,
if we have:

T()

m (m)
M N
TM T(N) T
m
(M) T
(m)
(N)
i.e., the following diagram commutes:
M m (m) N

T
m
M T
(m)
N

T
m

N
with the natural projection
M
: TM M, given by
M
(T
m
M) = m, that takes a
tangent vector v to the point m M at which the vector v is attached i.e., v T
m
M.
For an nD smooth manifold M, its nD tangent bundle TM is the disjoint union
of all its tangent spaces T
m
M at all points m M, TM =
_
mM
T
m
M.
To dene the smooth structure on TM, we need to specify how to construct
local coordinates on TM. To do this, let (x
1
(m), ..., x
n
(m)) be local coordinates of a
point m on M and let (v
1
(m), ..., v
n
(m)) be components of a tangent vector in this
coordinate system. Then the 2n numbers (x
1
(m), ..., x
n
(m), v
1
(m), ..., v
n
(m)) give
a local coordinate system on TM.
TM =
_
mM
T
m
M denes a family of vector spaces parameterized by M. The
inverse image
1
M
(m) of a point m M under the natural projection
M
is the
tangent space T
m
M. This space is called the bre of the tangent bundle over the
point m M.
16
A C
k
map : M N between two manifolds M and N induces a linear
tangent map T : TM TN between their tangent bundles, i.e., the following
diagram commutes:
M N

TM TN

N
All tangent bundles and their tangent maps form the category T B. The category
T B is the natural framework for Lagrangian dynamics.
Now, we can formulate the global version of the chain rule. If : M N and
: N P are two smooth maps, then we have T( ) = T T. In other
words, we have a functor T : /T B from the category / of smooth manifolds
to the category T B of their tangent bundles:
N P

( )
T
=

TN TP

T
TM
T

T( )

3.5 Cotangent Bundle and Hamiltonian Dynamics


The cotangent bundle of a smooth nmanifold is the place is where 1forms live, and
is itself a smooth 2nmanifold. Covectorelds (1forms) are cross-sections of the
cotangent bundle. The Hamiltonian is a natural energy function on the cotangent
bundle (see [8, 9]).
A dual notion to the tangent space T
m
M to a smooth manifold M at a point m
is its cotangent space T

m
M at the same point m. Similarly to the tangent bundle,
for a smooth manifold M of dimension n, its cotangent bundle T

M is the disjoint
union of all its cotangent spaces T

m
M at all points m M, i.e., T

M =
_
mM
T

m
M.
Therefore, the cotangent bundle of an nmanifold M is the vector bundle T

M =
(TM)

, the (real) dual of the tangent bundle TM.


If M is an nmanifold, then T

M is a 2nmanifold. To dene the smooth


structure on T

M, we need to specify how to construct local coordinates on T

M.
To do this, let (x
1
(m), ..., x
n
(m)) be local coordinates of a point m on M and let
(p
1
(m), ..., p
n
(m)) be components of a covector in this coordinate system. Then the
2n numbers (x
1
(m), ..., x
n
(m), p
1
(m), ..., p
n
(m)) give a local coordinate system on
T

M. This is the basic idea one uses to prove that indeed T

M is a 2nmanifold.
17
T

M =
_
mM
T

m
M denes a family of vector spaces parameterized by M, with
the conatural projection,

M
: T

M M, given by

M
(T

m
M) = m, that takes a
covector p to the point m M at which the covector p is attached i.e., p T

m
M.
The inverse image
1
M
(m) of a point m M under the conatural projection

M
is
the cotangent space T

m
M. This space is called the bre of the cotangent bundle
over the point m M.
In a similar way, a C
k
map : M N between two manifolds M and N induces
a linear cotangent map T

: T

M T

N between their cotangent bundles, i.e.,


the following diagram commutes:
M N

M T

N
All cotangent bundles and their cotangent maps form the category T

B. The
category T

B is the natural stage for Hamiltonian dynamics.


Now, we can formulate the dual version of the global chain rule. If : M N
and : N P are two smooth maps, then we have T

( ) = T

. In
other words, we have a cofunctor T

: /T

B from the category / of smooth


manifolds to the category T

B of their cotangent bundles:


N P

( )
T

N T

M
T

( )

4 Lie Groups
In this section we present the basics of classical theory of Lie groups and their Lie
algebras, as developed mainly by Sophus Lie, Elie Cartan, Felix Klein, Wilhelm
Killing and Hermann Weyl. For more comprehensive treatment see e.g., [21, 22, 23,
24, 25].
In the middle of the 19th Century S. Lie made a far reaching discovery that
techniques designed to solve particular unrelated types of ODEs, such as separable,
homogeneous and exact equations, were in fact all special cases of a general form of
integration procedure based on the invariance of the dierential equation under a
18
continuous group of symmetries. Roughly speaking a symmetry group of a system
of dierential equations is a group that transforms solutions of the system to other
solutions. Once the symmetry group has been identied a number of techniques
to solve and classify these dierential equations becomes possible. In the classical
framework of Lie, these groups were local groups and arose locally as groups of
transformations on some Euclidean space. The passage from the local Lie group to
the present day denition using manifolds was accomplished by E. Cartan at the
end of the 19th Century, whose work is a striking synthesis of Lie theory, classical
geometry, dierential geometry and topology.
These continuous groups, which originally appeared as symmetry groups of dif-
ferential equations, have over the years had a profound impact on diverse areas such
as algebraic topology, dierential geometry, numerical analysis, control theory, clas-
sical mechanics, quantum mechanics etc. They are now universally known as Lie
groups.
A Lie group is smooth manifold which also carries a group structure whose
product and inversion operations are smooth as maps of manifolds. These objects
arise naturally in describing physical symmetries.
6
6
Here are a few examples of Lie groups and their relations to other areas of mathematics and
physics:
1. Euclidean space R
n
is an Abelian Lie group (with ordinary vector addition as the group
operation).
2. The group GLn(R) of invertible matrices (under matrix multiplication) is a Lie group of
dimension n
2
. It has a subgroup SLn(R) of matrices of determinant 1 which is also a Lie
group.
3. The group On(R) generated by all rotations and reections of an nD vector space is a Lie
group called the orthogonal group. It has a subgroup of elements of determinant 1, called
the special orthogonal group SO(n), which is the group of rotations in R
n
.
4. Spin groups are double covers of the special orthogonal groups (used e.g., for studying
fermions in quantum eld theory).
5. The group Sp2n(R) of all matrices preserving a symplectic form is a Lie group called the
symplectic group.
6. The Lorentz group and the Poincare group of isometries of spacetime are Lie groups of
dimensions 6 and 10 that are used in special relativity.
7. The Heisenberg group is a Lie group of dimension 3, used in quantum mechanics.
8. The unitary group U(n) is a compact group of dimension n
2
consisting of unitary matrices.
It has a subgroup of elements of determinant 1, called the special unitary group SU(n).
9. The group U(1) SU(2) SU(3) is a Lie group of dimension 1 + 3 + 8 = 12 that is the
gauge group of the Standard Model of elementary particles, whose dimension corresponds to:
1 photon + 3 vector bosons + 8 gluons.
19
A Lie group is a group whose elements can be continuously parametrized by
real numbers, such as the rotation group SO(3), which can be parametrized by the
Euler angles. More formally, a Lie group is an analytic real or complex manifold
that is also a group, such that the group operations multiplication and inversion
are analytic maps. Lie groups are important in mathematical analysis, physics and
geometry because they serve to describe the symmetry of analytical structures. They
were introduced by Sophus Lie in 1870 in order to study symmetries of dierential
equations.
While the Euclidean space R
n
is a real Lie group (with ordinary vector addition
as the group operation), more typical examples are given by matrix Lie groups, i.e.,
groups of invertible matrices (under matrix multiplication). For instance, the group
SO(3) of all rotations in R
3
is a matrix Lie group.
One classies Lie groups regarding their algebraic properties
7
(simple, semisim-
ple, solvable, nilpotent, Abelian), their connectedness (connected or simply con-
nected) and their compactness.
8
To every Lie group, we can associate a Lie algebra which completely captures
the local structure of the group (at least if the Lie group is connected).
9
7
If G and H are Lie groups (both real or both complex), then a Liegrouphomomorphism
f : G H is a group homomorphism which is also an analytic map (one can show that it is
equivalent to require only that f be continuous). The composition of two such homomorphisms
is again a homomorphism, and the class of all (real or complex) Lie groups, together with these
morphisms, forms a category. The two Lie groups are called isomorphic i there exists a bijective
homomorphism between them whose inverse is also a homomorphism. Isomorphic Lie groups do
not need to be distinguished for all practical purposes; they only dier in the notation of their
elements.
8
An ntorus T
n
= S
1
S
1
S
1
(as dened above) is an example of a compact Abelian
Lie group. This follows from the fact that the unit circle S
1
is a compact Abelian Lie group (when
identied with the unit complex numbers with multiplication). Group multiplication on T
n
is then
dened by coordinatewise multiplication.
Toroidal groups play an important part in the theory of compact Lie groups. This is due in part
to the fact that in any compact Lie group one can always nd a maximal torus; that is, a closed
subgroup which is a torus of the largest possible dimension.
9
Conventionally, one can regard any eld X of tangent vectors on a Lie group as a partial
dierential operator, denoting by Xf the Lie derivative (the directional derivative) of the scalar
eld f in the direction of X. Then a vectoreld on a Lie group G is said to be leftinvariant if it
commutes with left translation, which means the following. Dene Lg[f](x) = f(gx) for any analytic
function f : G R and all g, x G. Then the vectoreld X is leftinvariant i XLg = LgX for
all g G. Similarly, instead of R, we can use C. The set of all vectorelds on an analytic manifold
is a Lie algebra over R (or C).
On a Lie group G, the leftinvariant vectorelds form a subalgebra, the Lie algebra g associated
with G. This Lie algebra is nitedimensional (it has the same dimension as the manifold G) which
makes it susceptible to classication attempts. By classifying g, one can also get a handle on the
group G. The representation theory of simple Lie groups is the best and most important example.
Every element v of the tangent space Te at the identity element e of G determines a unique
20
4.1 Denition of a Lie Group
A Lie group is a smooth (Banach) manifold M that has at the same time a group
Gstructure consistent with its manifold Mstructure in the sense that group mul-
tiplication : GG G, (g, h) gh and the group inversion : G G, g
g
1
are C
k
maps. A point e G is called the group identity element (see e.g.,
[21, 22, 1, 3]).
For example, any nD Banach vector space V is an Abelian Lie group with group
operations : V V V , (x, y) = x+y, and : V V , (x) = x. The identity
is just the zero vector. We call such a Lie group a vector group.
Let G and H be two Lie groups. A map G H is said to be a morphism of
Lie groups (or their smooth homomorphism) if it is their homomorphism as abstract
groups and their smooth map as manifolds.
Similarly, a group G which is at the same time a topological space is said to be
a topological group if both maps (, ) are continuous, i.e., C
0
maps for it. The
homomorphism G H of topological groups is said to be continuous if it is a
continuous map.
A topological group (as well as a smooth manifold) is not necessarily Hausdor.
leftinvariant vectoreld whose value at the element g of G is denoted by gv; the vector space
underlying the Lie algebra g may therefore be identied with Te.
Every vectoreld v in the Lie algebra g determines a function c : R G whose derivative
everywhere is given by the corresponding leftinvariant vectoreld: c

(t) = TL
c(t)
v and which
has the property: c(s +t) = c(s)c(t), (for all s and t) (the operation on the r.h.s. is the group
multiplication in G). The formal similarity of this formula with the one valid for the elementary
exponential function justies the denition: exp(v) = c(1). This is called the exponential map,
and it maps the Lie algebra g into the Lie group G. It provides a dieomorphism between a
neighborhood of 0 in g and a neighborhood of e in G. This exponential map is a generalization of
the exponential function for real numbers (since R is the Lie algebra of the Lie group of positive
real numbers with multiplication), for complex numbers (since C is the Lie algebra of the Lie
group of nonzero complex numbers with multiplication) and for matrices (since M(n, R) with the
regular commutator is the Lie algebra of the Lie group GL(n, R) of all invertible matrices). As the
exponential map is surjective on some neighborhood N of e, it is common to call elements of the
Lie algebra innitesimal generators of the group G.
The exponential map and the Lie algebra determine the local group structure of every connected
Lie group, because of the BakerCampbellHausdor formula: there exists a neighborhood U of
the zero element of the Lie algebra g, such that for u, v U we have
exp(u)exp(v) = exp(u +v + 1/2[u, v] + 1/12[[u, v], v] 1/12[[u, v], u] ...),
where the omitted terms are known and involve Lie brackets of four or more elements. In case u
and v commute, this formula reduces to the familiar exponential law:
exp(u)exp(v) = exp(u +v).
Every homomorphism f : G H of Lie groups induces a homomorphism between the corre-
sponding Lie algebras g and h. The association G = g is called the Lie Functor.
21
A topological group G is Hausdor i its identity is closed. As a corollary we have
that every Lie group is a Hausdor topological group.
For every g in a Lie group G, the two maps,
L
g
: G G, h gh,
R
h
: G G, g gh,
are called left and right translation maps. Since L
g
L
h
= L
gh
, and R
g
R
h
= R
gh
,
it follows that (L
g
)
1
= L
g
1 and (R
g
)
1
= R
g
1, so both L
g
and R
g
are dieomor-
phisms. Moreover L
g
R
h
= R
h
L
g
, i.e., left and right translation commute.
A vectoreld X on G is called leftinvariant vectoreld if for every g G,
L

g
X = X, that is, if (T
h
L
g
)X(h) = X(gh) for all h G, i.e., the following diagram
commutes:
G G

L
g
TG TG

TL
g

X
A Riemannian metric on a Lie group G is called left-invariant if it is preserved by
all left translations L
g
, i.e., if the derivative of left translation carries every vector to
a vector of the same length. Similarly, a vector eld X on G is called leftinvariant
if (for every g G) L

g
X = X.
4.2 Lie Algebra
An algebra A is a vector space with a product. The product must have the property
that
a(uv) = (au)v = u(av),
for every a R and u, v A. A map : A A

between algebras is called an


algebra homomorphism if (u v) = (u) (v). A vector subspace I of an algebra
A is called a left ideal (resp. right ideal ) if it is closed under algebra multiplication
and if u A and i I implies that ui I (resp. iu I). A subspace I is said
to be a twosided ideal if it is both a left and right ideal. An ideal may not be an
algebra itself, but the quotient of an algebra by a twosided ideal inherits an algebra
structure from A.
A Lie algebra is an algebra A where the multiplication, i.e., the Lie bracket
(u, v) [u, v], has the following properties:
LA 1. [u, u] = 0 for every u A, and
LA 2. [u, [v, w]] + [w, [u, v]] + [v, w, u]] = 0 for all u, v, w A.
22
The condition LA 2 is usually called Jacobi identity. A subspace E A of a Lie
algebra is called a Lie subalgebra if [u, v] E for every u, v E. A map : A A

between Lie algebras is called a Lie algebra homomorphism if ([u, v]) = [(u), (v)]
for each u, v A.
All Lie algebras (over a given eld K) and all smooth homomorphisms between
them form the category ///, which is itself a complete subcategory of the category
// of all algebras and their homomorphisms.
Let A
L
(G) denote the set of leftinvariant vectorelds on G; it is a Lie subalge-
bra of A(G), the set of all vectorelds on G, since L

g
[X, Y ] = [L

g
X, L

g
Y ] = [X, Y ],
so the Lie bracket [X, Y ] A
L
(G).
Let e be the identity element of G. Then for each on the tangent space T
e
G we
dene a vectoreld X

on G by X

(g) = T
e
L
g
(). A
L
(G) and T
e
G are isomorphic
as vector spaces. Dene the Lie bracket on T
e
G by [, ] = [X

, X

] (e) for all


, T
e
G. This makes T
e
G into a Lie algebra. Also, by construction, we have
[X

, X

] = X
[,]
; this denes a bracket in T
e
G via left extension. The vector space
T
e
G with the above algebra structure is called the Lie algebra of the Lie group G
and is denoted g.
For example, let V be a nD vector space. Then T
e
V V and the leftinvariant
vectoreld dened by T
e
V is the constant vectoreld X

() = , for all V .
The Lie algebra of V is V itself.
Since any two elements of an Abelian Lie group G commute, it follows that all
adjoint operators Ad
g
, g G, equal the identity. Therefore, the Lie algebra g is
Abelian; that is, [, ] = 0 for all , g.
For example, G = SO(3) is the group of rotations of 3D Euclidean space, i.e.
the conguration space of a rigid body xed at a point. A motion of the body is
then described by a curve g = g(t) in the group SO(3). Its Lie algebra g = so(3) is
the 3D vector space of angular velocities of all possible rotations. The commutator
in this algebra is the usual vector (cross) product (see, e.g. [1, 3, 9]).
A rotation velocity g of the rigid body (xed at a point) is a tangent vector to
the Lie group G = SO(3) at the point g G. To get the angular velocity, we must
carry this vector to the tangent space TG
e
of the group at the identity, i.e. to its Lie
algebra g = so(3). This can be done in two ways: by left and right translation, L
g
and R
g
. As a result, we obtain two dierent vector elds in the Lie algebra so(3) :

c
= L
g
1

g so(3) and
x
= R
g
1

g so(3),
which are called the angular velocity in the body and the angular velocity in space,
respectively.
The dual space g

to the Lie algebra g = so(3) is the space of angular momenta


. The kinetic energy T of a body is determined by the vector eld of angular
23
velocity in the body and does not depend on the position of the body in space.
Therefore, kinetic energy gives a left-invariant Riemannian metric on the rotation
group G = SO(3).
4.3 One-Parameter Subgroup
Let X

be a leftinvariant vectoreld on G corresponding to in g. Then there is


a unique integral curve

: R G of X

starting at e, i.e., (see, e.g. [8, 9])


(t) = X

(t)
_
,

(0) = e

(t) is a smooth oneparameter subgroup of G, i.e.,

(t +s) =

(t)

(s), since,
as functions of t both sides equal

(s) at t = 0 and both satisfy dierential equation


(t) = X

(t)
_
by left invariance of X

, so they are equal. Left invariance can be


also used to show that

(t) is dened for all t R. Moreover, if : R G is a


oneparameter subgroup of G, i.e., a smooth homomorphism of the additive group
R into G, then =

with =

(0), since taking derivative at s = 0 in the relation
(t +s) = (t) (s) gives

(t) = X

(0)
((t)) ,
so =

since both equal e at t = 0. Therefore, all oneparameter subgroups of G


are of the form

(t) for some g.


4.4 Exponential Map
The map exp : g G, given by (see, e.g. [3, 8, 9]):
exp() =

(1), exp(0) = e
is called the exponential map of the Lie algebra g of G into G. exp is a C
k
-
map, similar to the projection of tangent and cotangent bundles; exp is locally a
dieomorphism from a neighborhood of zero in g onto a neighborhood of e in G; if
f : G H is a smooth homomorphism of Lie groups, then
f exp
G
= exp
H
T
e
f .
Also, in this case
exp(s) =

(s).
Indeed, for xed s R, the curve t

(ts), which at t = 0 passes through e,


satises the dierential equation
d
dt

(ts) = sX

(ts)
_
= X
s
_

(ts)
_
.
24
Since
s
(t) satises the same dierential equation and passes through e at t = 0, it
follows that
s
(t) =

(st). Putting t = 1 induces exp(s) =

(s).
Hence exp maps the line s in g onto the oneparameter subgroup

(s) of G,
which is tangent to at e. It follows from left invariance that the ow F

t
of X
satises F

t
(g) = g exp(s).
Globally, the exponential map exp is a natural operation, i.e., for any morphism
: G H of Lie groups G and H and a Lie functor T, the following diagram
commutes:
G H

T(G) T(H)

T()

exp

exp
Let G
1
and G
2
be Lie groups with Lie algebras g
1
and g
2
. Then G
1
G
2
is a
Lie group with Lie algebra g
1
g
2
, and the exponential map is given by:
exp : g
1
g
2
G
1
G
2
, (
1
,
2
) (exp
1
(
1
), exp
2
(
2
)) .
For example, in case of a nD vector space, or innitedimensional Banach space,
the exponential map is the identity.
The unit circle in the complex plane S
1
= z C : [z[ = 1 is an Abelian Lie
group under multiplication. The tangent space T
e
S
1
is the imaginary axis, and we
identify R with T
e
S
1
by t 2it. With this identication, the exponential map
exp : R S
1
is given by exp(t) = e
2it
.
The nD torus T
n
= S
1
S
1
(n times) is an Abelian Lie group. The exponen-
tial map exp : R
n
T
n
is given by
exp(t
1
, ..., t
n
) = (e
2it
1
, ..., e
2itn
).
Since S
1
= R/Z, it follows that T
n
= R
n
/Z
n
, the projection R
n
T
n
being given
by the exp map.
4.5 Adjoint Representation
For every g G, the map (see, e.g. [1, 3, 8, 9]):
Ad
g
= T
e
_
R
g
1 L
g
_
: g g
is called the adjoint map, or adjoint operator associated with g.
For each g and g G we have
exp (Ad
g
) = g (exp ) g
1
.
25
The relation between the adjoint map and the Lie bracket is the following: For
all , g we have
d
dt

t=0
Ad
exp(t)
= [, ].
Left and right translations induce operators on the cotangent space T

G
g
dual
to L
g
and R
g
, denoted by (for every h G):
L

g
: T

G
gh
T

G
h
, R

g
: T

G
hg
T

G
h
.
The transpose operators Ad

g
: g g satisfy the relations Ad

gh
= Ad

h
Ad

g
(for every
g, h G) and constitute the co-adjoint representation of the Lie group G. The
co-adjoint representation plays an important role in all questions related to (left)
invariant metrics on the Lie group. According to A. Kirillov, the orbit of any vector
eld X in a Lie algebra g in a co-adjoint representation Ad

g
is itself a symplectic
manifold and therefore a phase space for a Hamiltonian mechanical system.
A Lie subgroup H of G is a subgroup H of G which is also a submanifold of
G. Then h is a Lie subalgebra of g and moreover h = g[ exp(t) H, for all
t R.
One can characterize Lebesgue measure up to a multiplicative constant on R
n
by
its invariance under translations. Similarly, on a locally compact group there is a
unique (up to a nonzero multiplicative constant) leftinvariant measure, called Haar
measure. For Lie groups the existence of such measures is especially simple: Let G be
a Lie group. Then there is a volume form Ub5, unique up to nonzero multiplicative
constants, that is leftinvariant. If G is compact, Ub5 is right invariant as well.
4.6 Actions of Lie Groups on Smooth Manifolds
Let M be a smooth manifold. An action of a Lie group G (with the unit element
e) on M is a smooth map : G M M, such that for all x M and g, h G,
(i) (e, x) = x and (ii) (g, (h, x)) = (gh, x). In other words, letting
g
: x
M
g
(x) = (g, x) M, we have (i)
e
= id
M
and (ii)
g

h
=
gh
.
g
is a
dieomorphism, since (
g
)
1
=
g
1. We say that the map g G
g
Diff(M)
is a homomorphism of G into the group of dieomorphisms of M. In case that M
is a vector space and each
g
is a linear operator, the function of G on M is called
a representation of G on M (see, e.g. [1, 3, 8, 9]).
An action of G on M is said to be transitive group action, if for every x, y M,
there is g G such that (g, x) = y; eective group action, if
g
= id
M
implies g = e,
that is g
g
is 11; and free group action, if for each x M, g
g
(x) is 11.
For example,
26
1. G = R acts on M = R by translations; explicitly,
: GM M, (s, x) = x +s.
Then for x R, O
x
= R. Hence M/G is a single point, and the action is
transitive and free.
2. A complete ow
t
of a vectoreld X on M gives an action of R on M, namely
(t, x) R M
t
(x) M.
3. Left translation L
g
: G G denes an eective action of G on itself. It is also
transitive.
4. The coadjoint action of G on g

is given by
Ad

: (g, ) Gg

Ad

g
1
() =
_
T
e
(R
g
1 L
g
)
_

.
Let be an action of G on M. For x M the orbit of x is dened by
O
x
=
g
(x)[g G M
and the isotropy group of at x is given by
G
x
= g G[(g, x) = x G.
An action of G on a manifold M denes an equivalence relation on M by the
relation belonging to the same orbit; explicitly, for x, y M, we write x y if there
exists a g G such that (g, x) = y, that is, if y O
x
. The set of all orbits M/G is
called the group orbit space (see, e.g. [1, 3, 8, 9]).
For example, let M = R
2
0, G = SO(2), the group of rotations in plane, and
the action of G on M given by
__
cos sin
sin cos
_
, (x, y)
_
(xcos y sin , xsin +y cos ).
The action is always free and eective, and the orbits are concentric circles, thus the
orbit space is M/G R

+
.
A crucial concept in mechanics is the innitesimal description of an action. Let
: GM M be an action of a Lie group G on a smooth manifold M. For each
g,

: R M M,

(t, x) = (exp(t), x)
27
is an R-action on M. Therefore,
exp(t)
: M M is a ow on M; the correspond-
ing vectoreld on M, given by

M
(x) =
d
dt

t=0

exp(t)
(x)
is called the innitesimal generator of the action, corresponding to in g.
The tangent space at x to an orbit O
x
is given by
T
x
O
x
=
M
(x)[ g.
Let : G M M be a smooth G action. For all g G, all , g and
all , R, we have:
(Ad
g
)
M
=

g
1

M
, [
M
,
M
] = [, ]
M
, and ( +)
M
=
M
+
M
.
Let M be a smooth manifold, G a Lie group and : G M M a Gaction
on M. We say that a smooth map f : M M is with respect to this action if for
all g G,
f
g
=
g
f.
Let f : M M be an equivariant smooth map. Then for any g we have
Tf
M
=
M
f.
4.7 Basic Tables of Lie Groups and Their Lie Algebras
One classies Lie groups regarding their algebraic properties (simple, semisimple,
solvable, nilpotent, Abelian), their connectedness (connected or simply connected)
and their compactness (see Tables A.1A.3). This is the content of the Hilbert 5th
problem.
28
Some real Lie groups and their Lie algebras:
Lie
group
Description Remarks Lie
algb.
Description dim
/R
R
n
Euclidean space
with addition
Abelian, simply
connected, not
compact
R
n
the Lie bracket
is zero
n
R

nonzero real
numbers with
multiplication
Abelian, not
connected, not
compact
R the Lie bracket
is zero
1
R
>0
positive real
numbers with
multiplication
Abelian, simply
connected, not
compact
R the Lie bracket
is zero
1
S
1
=
R/Z
complex num-
bers of absolute
value 1, with
multiplication
Abelian, con-
nected, not sim-
ply connected,
compact
R the Lie bracket
is zero
1
H

nonzero
quaternions
with multiplica-
tion
simply con-
nected, not
compact
H quaternions,
with Lie bracket
the commutator
4
S
3
quaternions of
absolute value
1, with mul-
tiplication; a
3sphere
simply con-
nected, com-
pact, simple
and semi
simple, isomor-
phic to SU(2),
SO(3) and to
Spin(3)
R
3
real 3vectors,
with Lie bracket
the cross prod-
uct; isomorphic
to su(2) and to
so(3)
3
GL(n, R) general linear
group: invert-
ible nby-n
real matrices
not connected,
not compact
M(n, R) nby-n matri-
ces, with Lie
bracket the
commutator
n
2
GL
+
(n, R) nby-n real
matrices with
positive deter-
minant
simply con-
nected, not
compact
M(n, R) nby-n matri-
ces, with Lie
bracket the
commutator
n
2
29
Classical real Lie groups and their Lie algebras:
Lie
group
Description Remarks Lie
algb.
Description dim
/R
SL(n, R) special linear
group: real
matrices with
determinant 1
simply con-
nected, not
compact if
n > 1
sl(n, R) square matrices
with trace 0,
with Lie bracket
the commutator
n
2

1
O(n, R) orthogonal
group: real
orthogonal
matrices
not connected,
compact
so(n, R) skew
symmetric
square real
matrices, with
Lie bracket
the commuta-
tor; so(3, R) is
isomorphic to
su(2) and to R
3
with the cross
product
n(n
1)/2
SO(n, R) special orthogo-
nal group: real
orthogonal ma-
trices with de-
terminant 1
connected, com-
pact, for n 2:
not simply con-
nected, for n =
3 and n
5: simple and
semisimple
so(n, R) skew
symmetric
square real
matrices, with
Lie bracket the
commutator
n(n
1)/2
Spin(n) spinor group simply con-
nected, com-
pact, for n = 3
and n 5:
simple and
semisimple
so(n, R) skew
symmetric
square real
matrices, with
Lie bracket the
commutator
n(n
1)/2
U(n) unitary group:
complex unitary
nby-n matri-
ces
isomorphic to
S
1
for n = 1,
not simply
connected,
compact
u(n) square com-
plex matrices
A satisfying
A = A

, with
Lie bracket the
commutator
n
2
SU(n) special unitary
group: complex
unitary nby-n
matrices with
determinant 1
simply con-
nected, com-
pact, for n 2:
simple and
semisimple
su(n) square complex
matrices A with
trace 0 satisfy-
ing A = A

,
with Lie bracket
the commutator
n
2

1
30
Basic complex Lie groups and their Lie algebras:
10
Lie
group
Description Remarks Lie
algb.
Description dim
/C
C
n
group operation
is addition
Abelian, simply
connected, not
compact
C
n
the Lie bracket
is zero
n
C

nonzero com-
plex numbers
with multiplica-
tion
Abelian, not
simply con-
nected, not
compact
C the Lie bracket
is zero
1
GL(n, C) general linear
group: invert-
ible nby-n
complex matri-
ces
simply con-
nected, not
compact, for
n = 1: iso-
morphic to
C

M(n, C) nby-n matri-


ces, with Lie
bracket the
commutator
n
2
SL(n, C) special linear
group: complex
matrices with
determinant 1
simple,
semisimple,
simply con-
nected, for
n 2: not
compact
sl(n, C) square matrices
with trace 0,
with Lie bracket
the commutator
n
2

1
O(n, C) orthogonal
group: com-
plex orthogonal
matrices
not connected,
for n 2: not
compact
so(n, C) skew
symmetric
square complex
matrices, with
Lie bracket the
commutator
n(n
1)/2
SO(n, C) special orthogo-
nal group: com-
plex orthogonal
matrices with
determinant 1
for n 2:
not compact,
not simply
connected, for
n = 3 and
n 5: simple
and semisimple
so(n, C) skew
symmetric
square complex
matrices, with
Lie bracket the
commutator
n(n
1)/2
4.8 Representations of Lie groups
The idea of a representation of a Lie group plays an important role in the study of
continuous symmetry (see, e.g., [22]). A great deal is known about such representa-
tions, a basic tool in their study being the use of the corresponding innitesimal
representations of Lie algebras.
Formally, a representation of a Lie group G on a vector space V (over a eld K)
10
The dimensions given are dimensions over C. Note that every complex Lie group/algebra can
also be viewed as a real Lie group/algebra of twice the dimension.
31
is a group homomorphism G Aut(V ) from G to the automorphism group of V .
If a basis for the vector space V is chosen, the representation can be expressed as a
homomorphism into GL(n, K). This is known as a matrix representation.
On the Lie algebra level, there is a corresponding linear map from the Lie algebra
of G to End(V ) preserving the Lie bracket [, ].
If the homomorphism is in fact an monomorphism, the representation is said to
be faithful.
A unitary representation is dened in the same way, except that G maps to
unitary matrices; the Lie algebra will then map to skewHermitian matrices.
Now, if G is a semisimple group, its nitedimensional representations can be
decomposed as direct sums of irreducible representations. The irreducibles are in-
dexed by highest weight; the allowable (dominant) highest weights satisfy a suitable
positivity condition. In particular, there exists a set of fundamental weights, indexed
by the vertices of the Dynkin diagram of G (see below), such that dominant weights
are simply nonnegative integer linear combinations of the fundamental weights.
If G is a commutative compact Lie group, then its irreducible representations
are simply the continuous characters of G. A quotient representation is a quotient
module of the group ring.
4.9 Root Systems and Dynkin Diagrams
A root system is a special conguration in Euclidean space that has turned out to
be fundamental in Lie theory as well as in its applications. Also, the classication
scheme for root systems, by Dynkin diagrams, occurs in parts of mathematics with
no overt connection to Lie groups (such as singularity theory, see e.g., [22]).
4.9.1 Denitions
Formally, a root system is a nite set of nonzero vectors (roots) spanning a
nitedimensional Euclidean space V and satisfying the following properties:
1. The only scalar multiples of a root in V which belong to are itself and
-.
2. For every root in V , the set is symmetric under reection through the
hyperplane of vectors perpendicular to .
3. If and are vectors in , the projection of 2 onto the line through is an
integer multiple of .
The rank of a root system is the dimension of V . Two root systems may
be combined by regarding the Euclidean spaces they span as mutually orthogonal
32
subspaces of a common Euclidean space. A root system which does not arise from
such a combination, such as the systems A
2
, B
2
, and G
2
in Figure 3, is said to be
irreducible.
Two irreducible root systems (V
1
,
1
) and (V
2
,
2
) are considered to be the same
if there is an invertible linear transformation V
1
V
2
which preserves distance up
to a scale factor and which sends
1
to
2
.
The group of isometries of V generated by reections through hyperplanes as-
sociated to the roots of is called the Weyl group of as it acts faithfully on the
nite set , the Weyl group is always nite.
4.9.2 Classication
It is not too dicult to classify the root systems of rank 2 (see Figure 3).
Figure 3: Classication of root systems of rank 2.
Whenever is a root system in V and W is a subspace of V spanned by =
W, then is a root system in W. Thus, our exhaustive list of root systems
of rank 2 shows the geometric possibilities for any two roots in a root system. In
particular, two such roots meet at an angle of 0, 30, 45, 60, 90, 120, 135, 150, or 180
degrees.
In general, irreducible root systems are specied by a family (indicated by a
letter A to G) and the rank (indicated by a subscript n). There are four innite
families:
A
n
(n 1), which corresponds to the special unitary group, SU(n + 1);
B
n
(n 2), which corresponds to the special orthogonal group, SO(2n + 1);
C
n
(n 3), which corresponds to the symplectic group, Sp(2n);
D
n
(n 4), which corresponds to the special orthogonal group, SO(2n),
as well as ve exceptional cases: E
6
, E
7
, E
8
, F
4
, G
2
.
33
4.9.3 Dynkin Diagrams
A Dynkin diagram is a graph with a few dierent kinds of possible edges (see Figure
4). The connected components of the graph correspond to the irreducible subalge-
bras of g. So a simple Lie algebras Dynkin diagram has only one component. The
rules are restrictive. In fact, there are only certain possibilities for each component,
corresponding to the classication of semisimple Lie algebras (see, e.g., [26]).
Figure 4: The problem of classifying irreducible root systems reduces to the problem
of classifying connected Dynkin diagrams.
The roots of a complex Lie algebra form a lattice of rank k in a Cartan subalgebra
h g, where k is the Lie algebra rank of g. Hence, the root lattice can be considered
a lattice in R
k
. A vertex, or node, in the Dynkin diagram is drawn for each Lie
algebra simple root, which corresponds to a generator of the root lattice. Between
two nodes and , an edge is drawn if the simple roots are not perpendicular. One
line is drawn if the angle between them is 2/3, two lines if the angle is 3/4, and
three lines are drawn if the angle is 5/6. There are no other possible angles between
Lie algebra simple roots. Alternatively, the number of lines N between the simple
roots and is given by
N = A

=
2 ,
[[
2
2 ,
[[
2
= 4 cos
2
,
where A

=
2,
||
2
is an entry in the Cartan matrix (A

) (for details on Cartan


matrix see, e.g., [22]). In a Dynkin diagram, an arrow is drawn from the longer root
to the shorter root (when the angle is 3/4 or 5/6).
Here are some properties of admissible Dynkin diagrams:
1. A diagram obtained by removing a node from an admissible diagram is admis-
sible.
2. An admissible diagram has no loops.
3. No node has more than three lines attached to it.
34
4. A sequence of nodes with only two single lines can be collapsed to give an
admissible diagram.
5. The only connected diagram with a triple line has two nodes.
A CoxeterDynkin diagram, also called a Coxeter graph, is the same as a Dynkin
diagram, but without the arrows. The Coxeter diagram is sucient to characterize
the algebra, as can be seen by enumerating connected diagrams.
The simplest way to recover a simple Lie algebra from its Dynkin diagram is to
rst reconstruct its Cartan matrix (A
ij
). The ith node and jth node are connected
by A
ij
A
ji
lines. Since A
ij
= 0 i A
ji
= 0, and otherwise A
ij
3, 2, 1, it is
easy to nd A
ij
and A
ji
, up to order, from their product. The arrow in the diagram
indicates which is larger. For example, if node 1 and node 2 have two lines between
them, from node 1 to node 2, then A
12
= 1 and A
21
= 2.
However, it is worth pointing out that each simple Lie algebra can be constructed
concretely. For instance, the innite families A
n
, B
n
, C
n
, and D
n
correspond to
the special linear Lie algebra gl(n + 1, C), the odd orthogonal Lie algebra so(2n +
1, C), the symplectic Lie algebra sp(2n, C), and the even orthogonal Lie algebra
so(2n, C). The other simple Lie algebras are called exceptional Lie algebras, and
have constructions related to the octonions.
To prove this classication Theorem, one uses the angles between pairs of roots to
encode the root system in a much simpler combinatorial object, the Dynkin diagram.
The Dynkin diagrams can then be classied according to the scheme given above.
To every root system is associated a corresponding Dynkin diagram. Otherwise,
the Dynkin diagram can be extracted from the root system by choosing a base, that
is a subset of which is a basis of V with the special property that every vector
in when written in the basis has either all coecients 0 or else all 0.
The vertices of the Dynkin diagram correspond to vectors in . An edge is
drawn between each nonorthogonal pair of vectors; it is a double edge if they make
an angle of 135 degrees, and a triple edge if they make an angle of 150 degrees. In
addition, double and triple edges are marked with an angle sign pointing toward the
shorter vector.
Although a given root system has more than one base, the Weyl group acts
transitively on the set of bases. Therefore, the root system determines the Dynkin
diagram. Given two root systems with the same Dynkin diagram, we can match up
roots, starting with the roots in the base, and show that the systems are in fact the
same.
Thus the problem of classifying root systems reduces to the problem of classifying
possible Dynkin diagrams, and the problem of classifying irreducible root systems
reduces to the problem of classifying connected Dynkin diagrams. Dynkin diagrams
35
encode the inner product on E in terms of the basis , and the condition that this
inner product must be positive denite turns out to be all that is needed to get the
desired classication (see Figure 4).
In detail, the individual root systems can be realized casebycase, as in the
following paragraphs:
A
n
. Let V be the subspace of R
n+1
for which the coordinates sum to 0, and
let be the set of vectors in V of length

2 and with integer coordinates in R


n+1
.
Such a vector must have all but two coordinates equal to 0, one coordinate equal to
1, and one equal to -1, so there are n
2
+n roots in all.
B
n
. Let V = R
n
, and let consist of all integer vectors in V of length 1 or

2.
The total number of roots is 2n
2
.
C
n
: Let V = R
n
, and let consist of all integer vectors in V of

2 together
with all vectors of the form 2, where is an integer vector of length 1. The total
number of roots is 2n
2
. The total number of roots is 2n
2
.
D
n
. Let V = R
n
, and let consist of all integer vectors in V of length

2. The
total number of roots is 2n(n 1).
E
n
. For V
8
, let V = R
8
, and let E
8
denote the set of vectors of length

2
such that the coordinates of 2 are all integers and are either all even or all odd.
Then E
7
can be constructed as the intersection of E
8
with the hyperplane of vectors
perpendicular to a xed root in E
8
, and E
6
can be constructed as the intersection
of E
8
with two such hyperplanes corresponding to roots and which are neither
orthogonal to one another nor scalar multiples of one another. The root systems
E
6
, E
7
, and E
8
have 72, 126, and 240 roots respectively.
F
4
. For F
4
, let V = R
4
, and let denote the set of vectors of length 1 or

2
such that the coordinates of 2 are all integers and are either all even or all odd.
There are 48 roots in this system.
G
2
. There are 12 roots in G
2
, which form the vertices of a hexagram.
4.9.4 Irreducible Root Systems
Irreducible root systems classify a number of related objects in Lie theory, notably:
1. Simple complex Lie algebras;
2. Simple complex Lie groups;
3. Simply connected complex Lie groups which are simple modulo centers; and
4. Simple compact Lie groups.
In each case, the roots are nonzero weights of the adjoint representation.
A root system can also be said to describe a plants root and associated systems.
36
4.10 Simple and Semisimple Lie Groups and Algebras
A simple Lie group is a Lie group which is also a simple group. These groups, and
groups closely related to them, include many of the socalled classical groups of
geometry, which lie behind projective geometry and other geometries derived from
it by the Erlangen programme of Felix Klein. They also include some exceptional
groups, that were rst discovered by those pursuing the classication of simple Lie
groups. The exceptional groups account for many special examples and congu-
rations in other branches of mathematics. In particular the classication of nite
simple groups depended on a thorough prior knowledge of the exceptional possi-
bilities.
The complete listing of the simple Lie groups is the basis for the theory of the
semisimple Lie groups and reductive groups, and their representation theory. This
has turned out not only to be a major extension of the theory of compact Lie groups
(and their representation theory), but to be of basic signicance in mathematical
physics.
Such groups are classied using the prior classication of the complex simple
Lie algebras. It has been shown that a simple Lie group has a simple Lie algebra
that will occur on the list given there, once it is complexied (that is, made into a
complex vector space rather than a real one). This reduces the classication to two
further matters.
The groups SO(p, q, R) and SO(p +q, R), for example, give rise to dierent real
Lie algebras, but having the same Dynkin diagram. In general there may be dierent
real forms of the same complex Lie algebra.
Secondly, the Lie algebra only determines uniquely the simply connected (uni-
versal) cover G

of the component containing the identity of a Lie group G. It may


well happen that G

is not actually a simple group, for example having a nontrivial


center. We have therefore to worry about the global topology, by computing the
fundamental group of G (an Abelian group: a Lie group is an Hspace). This was
done by Elie Cartan.
For an example, take the special orthogonal groups in even dimension. With
I a scalar matrix in the center, these are not actually simple groups; and having
a twofold spin cover, they arent simplyconnected either. They lie between G

and G, in the notation above.


Recall that a semisimple module is a module in which each submodule is a direct
summand. In particular, a semisimple representation is completely reducible, i.e.,
is a direct sum of irreducible representations (under a descending chain condition).
Similarly, one speaks of an Abelian category as being semisimple when every object
has the corresponding property. Also, a semisimple ring is one that is semisimple as
a module over itself.
37
A semisimple matrix is diagonalizable over any algebraically closed eld contain-
ing its entries. In practice this means that it has a diagonal matrix as its Jordan
normal form.
A Lie algebra g is called semisimple when it is a direct sum of simple Lie algebras,
i.e., nontrivial Lie algebras L whose only ideals are 0 and L itself. An equivalent
condition is that the Killing form
B(X, Y ) = Tr(Ad(X) Ad(Y ))
is nondegenerate [27]. The following properties can be proved equivalent for a
nitedimensional algebra L over a eld of characteristic 0:
1. L is semisimple.
2. L has no nonzero Abelian ideal.
3. L has zero radical (the radical is the biggest solvable ideal).
4. Every representation of L is fully reducible, i.e., is a sum of irreducible repre-
sentations.
5. L is a (nite) direct product of simple Lie algebras (a Lie algebra is called
simple if it is not Abelian and has no nonzero ideal ).
A connected Lie group is called semisimple when its Lie algebra is semisimple;
and the same holds for algebraic groups. Every nite dimensional representation
of a semisimple Lie algebra, Lie group, or algebraic group in characteristic 0 is
semisimple, i.e., completely reducible, but the converse is not true. Moreover, in
characteristic p > 0, semisimple Lie groups and Lie algebras have nite dimensional
representations which are not semisimple. An element of a semisimple Lie group or
Lie algebra is itself semisimple if its image in every nitedimensional representation
is semisimple in the sense of matrices.
Every semisimple Lie algebra g can be classied by its Dynkin diagram [22].
5 Some Classical Examples of Lie Groups
5.1 Galilei Group
The Galilei group is the group of transformations in space and time that connect
those Cartesian systems that are termed inertial frames in Newtonian mechanics.
The most general relationship between two such frames is the following. The origin
of the time scale in the inertial frame S

may be shifted compared with that in S;


the orientation of the Cartesian axes in S

may be dierent from that in S; the


origin O of the Cartesian frame in S

may be moving relative to the origin O in S


at a uniform velocity. The transition from S to S

involves ten parameters; thus the


Galilei group is a ten parameter group. The basic assumption inherent in Galilei
Newtonian relativity is that there is an absolute time scale, so that the only way
38
in which the time variables used by two dierent inertial observers could possibly
dier is that the zero of time for one of them may be shifted relative to the zero of
time for the other (see, e.g. [1, 8, 9]).
Galilei spacetime structure involves the following three elements:
1. World, as a 4D ane space A
4
. The points of A
4
are called world points or
events. The parallel transitions of the world A
4
form a linear (i.e., Euclidean)
space R
4
.
2. Time, as a linear map t : R
4
R of the linear space of the world parallel
transitions onto the real time axes. Time interval from the event a A
4
to
b A
4
is called the number t(b a); if t(b a) = 0 then the events a and b are
called synchronous. The set of all mutually synchronous events consists a 3D
ane space A
3
, being a subspace of the world A
4
. The kernel of the mapping
t consists of the parallel transitions of A
4
translating arbitrary (and every)
event to the synchronous one; it is a linear 3D subspace R
3
of the space R
4
.
3. Distance (metric) between the synchronous events,
(a, b) =| a b |, for all a, b A
3
,
given by the scalar product in R
3
. The distance transforms arbitrary space of
synchronous events into the well known 3D Euclidean space E
3
.
The space A
4
, with the Galilei spacetime structure on it, is called Galilei space.
Galilei group is the group of all possible transformations of the Galilei space, pre-
serving its structure. The elements of the Galilei group are called Galilei transfor-
mations. Therefore, Galilei transformations are ane transformations of the world
A
4
preserving the time intervals and distances between the synchronous events.
The direct product R R
3
, of the time axes with the 3D linear space R3 with
a xed Euclidean structure, has a natural Galilei structure. It is called Galilei
coordinate system.
5.2 General Linear Group
The group of linear isomorphisms of R
n
to R
n
is a Lie group of dimension n
2
, called
the general linear group and denoted Gl(n, R). It is a smooth manifold, since it is
a subset of the vector space L(R
n
, R
n
) of all linear maps of R
n
to R
n
, as Gl(n, R)
is the inverse image of R0 under the continuous map A det A of L(R
n
, R
n
) to
R. The group operation is composition (see, e.g. [1, 3, 8, 9]).
(A, B) Gl(n, R) Gl(n, R) A B Gl(n, R)
39
and the inverse map is
A Gl(n, R) A
1
Gl(n, R).
If we choose a basis in R
n
, we can represent each element A Gl(n, R) by an
invertible (n n)-matrix. The group operation is then matrix multiplication and
the inversion is matrix inversion. The identity is the identity matrix I
n
. The group
operations are smooth since the formulas for the product and inverse of matrices are
smooth in the matrix components.
The Lie algebra of Gl(n, R) is gl(n), the vector space L(R
n
, R
n
) of all linear
transformations of R
n
, with the commutator bracket
[A, B] = AB BA.
For every A L(R
n
, R
n
),

A
: t R
A
(t) =

i=0
t
i
i!
A
i
Gl(n, R)
is a oneparameter subgroup of Gl(n, R), because

A
(0) = I and
A
(t) =

i=0
t
i1
(i 1)!
A
i
=
A
(t) A
Hence
A
is an integral curve of the leftinvariant vectoreld X
A
. Therefore, the
exponential map is given by
exp : A L(R
n
, R
n
) exp(A) e
A
=
A
(1) =

i=0
A
i
i!
Gl(n, R).
For each A Gl(n, R) the corresponding adjoint map
Ad
A
: L(R
n
, R
n
) L(R
n
, R
n
)
is given by
Ad
A
B = A B A
1
.
5.3 Rotational Lie Groups in Human/Humanoid Biomechanics
Local kinematics at each rotational robot or (synovial) human joint, is dened as
a group action of an nD constrained rotational Lie group SO(n) on the Euclidean
space R
n
. In particular, there is an action of SO(2)-group in uniaxial human joints
40
(cylindrical, or hinge joints, like knee and elbow) and an action of SO(3)-group
in threeaxial human joints (spherical, or ballandsocket joints, like hip, shoulder,
neck, wrist and ankle). In both cases, SO(n) acts, with its operators of rotation,
on the vector x = x

, (i = 1, 2, 3) of external, Cartesian coordinates of the parent


bodysegment, depending, at the same time, on the vector q = q
s
, (s = 1, , n)
on n groupparameters, i.e., joint angles (see [5, 6, 8, 9]).
Each joint rotation R SO(n) denes a map
R : x

, R(x

, q
s
) = R
q
sx

,
where R
q
s SO(n) are joint group operators. The vector v = v
s
, (s = 1, , n)
of n innitesimal generators of these rotations, i.e., joint angular velocities, given by
v
s
=
_
R(x

, q
s
)
q
s
_
q=0

constitute an nD Lie algebra so(n) corresponding to the joint rotation group SO(n).
Conversely, each joint group operator R
q
s, representing a oneparameter subgroup of
SO(n), is dened as the exponential map of the corresponding joint group generator
v
s
R
q
s = exp(q
s
v
s
)
This exponential map represents a solution of the joint operator dierential equation
in the joint groupparameter space q
s

dR
q
s
dq
s
= v
s
R
q
s.
5.3.1 Uniaxial Group of Joint Rotations
The uniaxial joint rotation in a single Cartesian plane around a perpendicular axis,
e.g., xyplane about the z axis, by an internal joint angle , leads to the following
transformation of the joint coordinates:
x

= xcos y sin , y

= xsin +y cos .
In this way, the joint SO(2)group, given by
SO(2) =
_
R

=
_
cos sin
sin cos
_
[ [0, 2]
_
,
acts in a canonical way on the Euclidean plane R
2
by
SO(2) =
__
cos sin
sin cos
_
,
_
x
y
__

_
xcos y sin
xsin y cos
_
.
41
Its associated Lie algebra so(2) is given by
so(2) =
__
0 t
t 0
_
[t R
_
,
since the curve

SO(2) given by

: t R

(t) =
_
cos t sint
sint cos t
_
SO(2),
passes through the identity I
2
=
_
1 0
0 1
_
and then
d
dt

t=0

(t) =
_
0
0
_
,
so that I
2
is a basis of so(2), since dim(SO(2)) = 1.
The exponential map exp : so(2) SO(2) is given by
exp
_
0
0
_
=

(1) =
_
cos t sint
sin t cos t
_
.
The innitesimal generator of the action of SO(2) on R
2
, i.e., joint angular
velocity v, is given by
v = y

x
+x

y
,
since
v
R
2 (x, y) =
d
dt

t=0
exp(tv) (x, y) =
d
dt

t=0
_
cos tv sin tv
sin tv cos tv
__
x
y
_
.
The momentum map J : T

R
2
R associated to the lifted action of SO(2) on
T

R
2
R
4
is given by
J (x, y, p
1
, p
2
) = xp
y
yp
x
, since
J (x, y, p
x
, p
y
) () = (p
x
dx +p
y
dy)(v
R
2) = vp
x
y +vp
y
x.
The Lie group SO(2) acts on the symplectic manifold (R
4
, = dp
x
dx+dp
y
dx)
by

__
cos sin
sin cos
_
, (x, y, p
x
, p
y
)
_
= (xcos y sin, xsin +y cos , p
x
cos p
y
sin , p
x
sin +p
y
cos ) .
42
5.3.2 ThreeAxial Group of Joint Rotations
The threeaxial SO(3)group of humanlike joint rotations depends on three pa-
rameters, Euler joint angles q
i
= (, , ), dening the rotations about the Cartesian
coordinate triedar (x, y, z) placed at the joint pivot point. Each of the Euler angles
are dened in the constrained range (, ), so the joint group space is a constrained
sphere of radius (see [5, 6, 8, 9]).
Let G = SO(3) = A /
33
(R) : A
t
A = I
3
, det(A) = 1 be the group
of rotations in R
3
. It is a Lie group and dim(G) = 3. Let us isolate its one
parameter joint subgroups, i.e., consider the three operators of the nite joint rota-
tions R

, R

, R

SO(3), given by
R

=
_
_
1 0 0
0 cos sin
0 sin cos
_
_
, R

=
_
_
cos 0 sin
0 1 0
sin 0 cos
_
_
, R

=
_
_
cos sin 0
sin cos 0
0 0 1
_
_
corresponding respectively to rotations about xaxis by an angle , about yaxis
by an angle , and about zaxis by an angle .
The total threeaxial joint rotation A is dened as the product of above one
parameter rotations R

, R

, R

, i.e., A = R

is equal
11
A =
_
_
cos cos cos sin sin cos cos + cos cos sin sin sin
sin cos cos sin sin sin sin + cos cos cos sin cos
sin sin sin cos cos
_
_
.
However, the order of these matrix products matters: dierent order products give
dierent results, as the matrix product is noncommutative product. This is the
reason why Hamiltons quaternions
12
are today commonly used to parameterize the
SO(3)group, especially in the eld of 3D computer graphics.
The oneparameter rotations R

, R

, R

dene curves in SO(3) starting from


I
3
=
_
1 0 0
0 1 0
0 0 1
_
. Their derivatives in = 0, = 0 and = 0 belong to the asso-
ciated tangent Lie algebra so(3). That is the corresponding innitesimal generators
of joint rotations joint angular velocities v

, v

, v

so(3) are respectively given


11
Note that this product is noncommutative, so it really depends on the order of multiplications.
12
Recall that the set of Hamiltons quaternions H represents an extension of the set of complex
numbers C. We can compute a rotation about the unit vector, u by an angle . The quaternion q
that computes this rotation is
q =
_
cos

2
, usin

2
_
.
43
by
v

=
_
0 0 0
0 0 1
0 1 0
_
= y

z
+z

y
, v

=
_
0 0 1
0 0 0
1 0 0
_
= z

x
+x

z
,
v

=
_
0 1 0
1 1 0
0 0 0
_
= x

y
+y

x
.
Moreover, the elements are linearly independent and so
so(3) =
_
_
_
_
_
0 a b
a 0
b 0
_
_
[a, b, R
_
_
_
.
The Lie algebra so(3) is identied with R
3
by associating to each v = (v

, v

, v

)
R
3
the matrix v so(3) given by v =
_
0 a b
a 0
b 0
_
. Then we have the following
identities:
1.

u v = [ u, v]; and
2. u v =
1
2
Tr( u v).
The exponential map exp : so(3) SO(3) is given by Rodrigues relation
exp(v) = I +
sin |v|
|v|
v +
1
2
_
sin
v
2
v
2
_
2
v
2
where the norm |v| is given by
|v| =
_
(v
1
)
2
+ (v
2
)
2
+ (v
3
)
2
.
The the dual, cotangent Lie algebra so(3)

, includes the three joint angular mo-


menta p

, p

, p

so(3)

, derived from the joint velocities v by multiplying them


with corresponding moments of inertia.
Note that the parameterization of SO(3)rotations is the subject of continuous
research and development in many theoretical and applied elds of mechanics, such
as rigid body, structural, and multibody dynamics, robotics, spacecraft attitude
dynamics, navigation, image processing, etc.
44
5.3.3 The Heavy Top
Consider a rigid body moving with a xed point but under the inuence of gravity.
This problem still has a conguration space SO(3), but the symmetry group is only
the circle group S
1
, consisting of rotations about the direction of gravity. One says
that gravity has broken the symmetry from SO(3) to S
1
. This time, eliminating the
S
1
symmetry mysteriously leads one to the larger Euclidean group SE(3) of rigid
motion of R
3
. Conversely, we can start with SE(3) as the conguration space for
the rigidbody and reduce out translations to arrive at SO(3) as the conguration
space. The equations of motion for a rigid body with a xed point in a gravitational
eld give an interesting example of a system that is Hamiltonian. The underlying
Lie algebra consists of the algebra of innitesimal Euclidean motions in R
3
(see
[1, 3, 8, 9]).
The basic phasespace we start with is again T

SO(3), parameterized by Euler


angles and their conjugate momenta. In these variables, the equations are in canon-
ical Hamiltonian form. However, the presence of gravity breaks the symmetry, and
the system is no longer SO(3) invariant, so it cannot be written entirely in terms of
the body angular momentum p. One also needs to keep track of , the direction of
gravity as seen from the body. This is dened by = A
1
k, where k points upward
and A is the element of SO(3) describing the current conguration of the body. The
equations of motion are
p
1
=
I
2
I
3
I
2
I
3
p
2
p
3
+Mgl(
2

2
),
p
2
=
I
3
I
1
I
3
I
1
p
3
p
1
+Mgl(
3

3
),
p
3
=
I
1
I
2
I
1
I
2
p
1
p
2
+Mgl(
1

1
),
and

= ,
where is the bodys angular velocity vector, I
1
, I
2
, I
3
are the bodys principal
moments of inertia, M is the bodys mass, g is the acceleration of gravity, is the
body xed unit vector on the line segment connecting the xed point with the bodys
center of mass, and l is the length of this segment.
5.4 Euclidean Groups of Rigid Body Motion
In this subsection we give description of two most important Lie groups in classical
mechanics in 2D and 3D, SE(2) and SE(3), respectively (see [4, 8, 9]).
45
5.4.1 Special Euclidean Group SE(2) in the Plane
The motion in uniaxial human joints is naturally modelled by the special Euclidean
group in the plane, SE(2). It consists of all transformations of R
2
of the form Az+a,
where z, a R
2
, and
A SO(2) =
_
matrices of the form
_
cos sin
sin cos
__
.
In other words, group SE(2) consists of matrices of the form:
(R

, a) =
_
R

a
0 I
_
, where a R
2
and R

is the rotation matrix:


R

=
_
cos sin
sin cos
_
, while I is the 3 3 identity matrix. The inverse (R

, a)
1
is given by
(R

, a)
1
=
_
R

a
0 I
_
1
=
_
R

a
0 I
_
.
The Lie algebra se(2) of SE(2) consists of 3 3 block matrices of the form
_
J v
0 0
_
, where J =
_
0 1
1 0
_
, (J
T
= J
1
= J),
with the usual commutator bracket. If we identify se(2) with R
3
by the isomorphism
_
J v
0 0
_
se(2) (, v) R
3
,
then the expression for the Lie algebra bracket becomes
[(, v
1
, v
2
), (, w
1
, w
2
)] = (0, v
2
w
2
, w
1
v1) = (0, J
T
w J
T
v),
where v = (v
1
, v
2
) and w = (w
1
, w
2
).
The adjoint group action of
(R

, a)
_
R

a
0 I
_
on (, v) =
_
J v
0 0
_
is given by the group conjugation,
_
R

a
0 I
__
J v
0 0
__
R

a
0 I
_
=
_
J Ja +R

v
0 0
_
,
46
or, in coordinates,
Ad
(R

,a)
(, v) = (, Ja +R

v). (2)
In proving (2) we used the identity R

J = JR

. Identify the dual algebra, se(2)

,
with matrices of the form
_

2
J 0
0
_
, via the nondegenerate pairing given by the
trace of the product. Thus, se(2)

is isomorphic to R
3
via
_

2
J 0
0
_
se(2)

(, ) R
3
,
so that in these coordinates, the pairing between se(2)

and se(2) becomes


(, ), (, v) = + v,
that is, the usual dot product in R
3
. The coadjoint group action is thus given by
Ad

(R

,a)
1
(, ) = ( R

Ja +R

). (3)
Formula (3) shows that the coadjoint orbits are the cylinders T

S
1

= (, )[ || =
const if ,= 0 together with the points are on the axis. The canonical cotangent
bundle projection : T

S
1

S
1

is dened as (, ) = .
5.4.2 Special Euclidean Group SE(3) in the 3D Space
The most common group structure in human biodynamics is the special Euclidean
group in 3D space, SE(3). Briey, the Euclidean SE(3)group is dened as a
semidirect (noncommutative) product of 3D rotations and 3D translations, SE(3) :=
SO(3) R
3
(see [4, 8, 9]). Its most important subgroups are the following:
Subgroup Denition
SO(3), group of rotations
in 3D (a spherical joint)
Set of all proper orthogonal
3 3 rotational matrices
SE(2), special Euclidean group
in 2D (all planar motions)
Set of all 3 3 matrices:
_
_
cos sin r
x
sin cos r
y
0 0 1
_
_
SO(2), group of rotations in 2D
subgroup of SE(2)group
(a revolute joint)
Set of all proper orthogonal
2 2 rotational matrices
included in SE(2) group
R
3
, group of translations in 3D
(all spatial displacements)
Euclidean 3D vector space
47
Lie Group SE(3) and Its Lie Algebra An element of SE(3) is a pair (A, a)
where A SO(3) and a R
3
. The action of SE(3) on R
3
is the rotation A followed
by translation by the vector a and has the expression
(A, a) x = Ax +a.
Using this formula, one sees that multiplication and inversion in SE(3) are given by
(A, a)(B, b) = (AB, Ab +a) and (A, a)
1
= (A
1
, A
1
a),
for A, B SO(3) and a, b R
3
. The identity element is (l, 0).
The Lie algebra of the Euclidean group SE(3) is se(3) = R
3
R
3
with the Lie
bracket
[(, u), (, v)] = ( , v u). (4)
The Lie algebra of the Euclidean group has a structure that is a special case of
what is called a semidirect product. Here it is the product of the group of rotations
with the corresponding group of translations. It turns out that semidirect products
occur under rather general circumstances when the symmetry in T

G is broken (see
[4, 8, 9]).
The dual Lie algebra of the Euclidean group SE(3) is se(3)

= R
3
R
3
with the
same Lie bracket (5).
Representation of SE(3) In other words, SE(3) := SO(3) R
3
is the Lie group
consisting of isometries of R
3
.
Using homogeneous coordinates, we can represent SE(3) as follows,
SE(3) =
__
R p
0 1
_
GL(4, R) : R SO(3), p R
3
_
,
with the action on R
3
given by the usual matrixvector product when we identify
R
3
with the section R
3
1 R
4
. In particular, given
g =
_
R p
0 1
_
SE(3),
and q R
3
, we have
g q = Rq +p,
or as a matrixvector product,
_
R p
0 1
__
q
1
_
=
_
Rq +p
1
_
.
48
Lie algebra of SE(3) The Lie algebra of SE(3) is given by
se(3) =
__
v
0 0
_
M
4
(R) : so(3), v R
3
_
,
where the attitude matrix : R
3
so(3) is given by
=
_
_
0
z

y

z
0
x

y

x
0
_
_
.
The exponential map of SE(3) The exponential map, exp : se(3) SE(3), is
given by
exp
_
v
0 0
_
=
_
exp() Av
0 1
_
,
where
A = I +
1 cos ||
||
2
+
|| sin ||
||
3

2
,
and exp() is given by the Rodriguez formula,
exp() = I +
sin ||
||
+
1 cos ||
||
2

2
.
In other words, the special Euclidean group SE(3) := SO(3) R
3
is the Lie
group consisting of isometries of the Euclidean 3D space R
3
. An element of SE(3)
is a pair (A, a) where A SO(3) and a R
3
. The action of SE(3) on R
3
is the
rotation A followed by translation by the vector a and has the expression
(A, a) x = Ax +a.
The Lie algebra of the Euclidean group SE(3) is se(3) = R
3
R
3
with the Lie
bracket
[(, u), (, v)] = ( , v u). (5)
Using homogeneous coordinates, we can represent SE(3) as follows,
SE(3) =
__
R p
0 1
_
GL(4, R) : R SO(3), p R
3
_
,
with the action on R
3
given by the usual matrixvector product when we identify
R
3
with the section R
3
1 R
4
. In particular, given
g =
_
R p
0 1
_
SE(3),
49
and q R
3
, we have
g q = Rq +p,
or as a matrixvector product,
_
R p
0 1
__
q
1
_
=
_
Rq +p
1
_
.
The Lie algebra of SE(3), denoted se(3), is given by
se(3) =
__
v
0 0
_
M
4
(R) : so(3), v R
3
_
,
where the attitude (or, angular velocity) matrix : R
3
so(3) is given by
=
_
_
0
z

y

z
0
x

y

x
0
_
_
.
The exponential map, exp : se(3) SE(3), is given by
exp
_
v
0 0
_
=
_
exp() Av
0 1
_
,
where
A = I +
1 cos ||
||
2
+
|| sin ||
||
3

2
,
and exp() is given by the Rodriguez formula,
exp() = I +
sin ||
||
+
1 cos ||
||
2

2
.
5.5 Basic Mechanical Examples
5.5.1 SE(2)Hovercraft
Conguration manifold is (, x, y) SE(2), given by matrix
P =
_
_
cos sin x
sin cos y
0 0 1
_
_
.
50
Kinematic equations of motion in Lie algebra se(2):

P = P
_
_
0 v
x
0 v
y
0 0 0
_
_
, ( =

, v
x
= x, v
y
= y).
Kinetic energy:
E
k
=
1
2
m(v
2
x
+v
2
y
) +
1
2
I
2
,
where m, I are mass and inertia moment of the hovercraft.
Dynamical equations of motion:
m v
x
= mv
y
+u
1
,
m v
y
= mv
x
+u
2
,
I = u
2
,
where = h is the torque applied at distance h from the centerofmass, while
u
1
, u
2
are control inputs.
5.5.2 SO(3)Satellite
Conguration manifold is rotation matrix R SO(3), with associated angular
velocity (attitude) matrix = (
1
,
2
,
3
) so(3) R
3
given by
so(3)
_
_
0
3

2

3
0
1

2

1
0
_
_
.
Kinematic equation of motion in so(3):

R = R,
Kinetic energy:
E
k
=
1
2

T
I,
where inertia tensor I is given by diagonal matrix,
I = diagI
1
, I
2
, I
3
.
Dynamical Euler equations of motion:
I = I +
i
u
i
,
where is the crossproduct in 3D,
i
are three external torques and u
i
= u
i
(t) are
control inputs.
51
5.5.3 SE(3)Submarine
The motion of a rigid body in incompressible, irrotational and inviscid uid is de-
ned by the conguration manifold SE(3), given by a pair of rotation matrix and
translation vector, (R, p) SE(3), such that angular velocity (attitude) matrix and
linear velocity vector, (, v) se(3) R
6
.
Kinematic equations of motion in se(3):
p = Rv,

R = R.
Kinetic energy (symmetrical):
E
k
=
1
2
v
T
Mv +
1
2

T
I,
where mass and inertia matrices are diagonal (for a neutrally buoyant ellipsoidal
body with uniformly distributed mass),
M = diagm
1
, m
2
, m
3
,
I = diagI
1
, I
2
, I
3
.
Dynamical Kirchho equations of motion read:
M v = Mv , I = I +Mv v.
By including the bodyxed external forces and torques, f
i
,
i
, with input controls
u
i
= u
(
t), the dynamical equations become:
M v = Mv +f
i
u
i
,
I = I +Mv v +
i
u
i
.
5.6 NewtonEuler SE(3)Dynamics
5.6.1 SO(3) : Euler Equations of Rigid Rotations
Unforced Euler equations read in vector form
I = , with I = diagI
1
, I
2
, I
3

and in scalar form


I
1

1
= (I
2
I
3
)
2

3
I
2

2
= (I
3
I
1
)
3

1
I
3

3
= (I
1
I
2
)
1

2
.
52
Using rotational kineticenergy Lagrangian
L() = E
rot
k
=
1
2

t
I =
1
2
(I
1

2
1
+I
2

2
2
+I
3

2
3
) (
t
= transpose)
Regarding the angular momentum =

L = I = (I
1

1
, I
2

2
, I
3

3
) as a vector,
we can derive unforced Euler equations: = as a system of EulerLagrange
Kirchho equations
d
dt

L =

L .
Forced Euler equations read in vector form
+ = T
and in scalar form
I
1

1
+ (I
3
I
2
)
2

3
= T
1
I
2

2
+ (I
1
I
3
)
3

1
= T
2
I
3

3
+ (I
2
I
1
)
1

2
= T
3
5.6.2 SE(3) : Coupled NewtonEuler Equations
Forced coupled NewtonEuler equations read in vector form
p M v = F +p , with M = diagm
1
, m
2
, m
3

I = T+ +p v, I = diagI
1
, I
2
, I
3
,
with principal inertia moments given in Cartesian coordinates (x, y, z) by density
dependent volume integrals
I
1
=
___
(z
2
+y
2
)dxdydz, I
2
=
___
(x
2
+y
2
)dxdydz, I
3
=
___
(x
2
+y
2
)dxdydz,
In tensor form, the forcedcoupled NewtonEuler equations read
p
i
M
ij
v
j
= F
i
+
j
ik
p
j

k
,

i
I
ij

j
= T
i
+
j
ik

k
+
j
ik
p
j
v
k
,
where the permutation symbol
j
ik
is dened as

j
ik
=
_

_
+1 if (i, j, k) is (1, 2, 3), (3, 1, 2) or (2, 3, 1),
1 if (i, j, k) is (3, 2, 1), (1, 3, 2) or (2, 1, 3),
0 otherwise: i = j or j = k or k = i.
53
In scalar form these equations read
p
1
= F
1
m
3
v
3

2
+m
2
v
2

3
p
2
= F
2
+m
3
v
3

1
m
1
v
1

3
p
3
= F
3
m
2
v
2

1
+m
1
v
1

2

1
= T
1
+ (m
2
m
3
)v
2
v
3
+ (I
2
I
3
)
2

3

2
= T
2
+ (m
3
m
1
)v
1
v
3
+ (I
3
I
1
)
1

3

3
= T
3
+ (m
1
m
2
)v
1
v
2
+ (I
1
I
2
)
1

2
.
These coupled rigidbody equations can be derived from the NewtonEuler kinetic
energy
E
k
=
1
2
v
t
Mv +
1
2

t
I
or, in tensor form
E =
1
2
M
ij
v
i
v
j
+
1
2
I
ij

i

j
.
Using the KirchhoLagrangian equations
d
dt

v
E
k
=
v
E
k
+F
d
dt

E
k
=

E
k
+
v
E
k
v +T,
or, in tensor form
d
dt

v
i E =
j
ik
(
v
j E)
k
+F
i
,
d
dt

i E =
j
ik
(

j E)
k
+
j
ik
(
v
j E) v
k
+T
i
we can derive linear and angular momentum covectors
p =
v
E
k
, =

E
k
or, in tensor form
p
i
=
v
i E,
i
=

i E,
and in scalar form
p = [p
1
, p
2
, p
3
] = [m
1
v
1
, m
2
v
2
, m
2
v
3
]
= [
1
,
2
,
3
] = [I
1

1
, I
2

2
, I
3

3
],
54
with their respective time derivatives, in vector form
p =
d
dt
p =
d
dt

v
E
k
, =
d
dt
=
d
dt

E
k
or, in tensor form
p
i
=
d
dt
p
i
=
d
dt

v
i E,
i
=
d
dt

i
=
d
dt

i E,
and in scalar form
p = [ p
1
, p
2
, p
3
] = [m
1
v
1
, m
2
v
2
, m
3
v
3
]
= [
1
,
2
,
3
] = [I
1

1
, I
2

2
, I
3

3
].
In addition, for the purpose of biomechanical injury prediction/prevention, we
have linear and angular jolts, respectively given in vector form by

F = p p p

T = p v p v,
or, in tensor form
13

F
i
= p
i

j
ik
p
j

j
ik
p
j

k
,

T
i
=
i

j
ik

j

j
ik

j

k

j
ik
p
j
v
k

j
ik
p
j
v
k
,
where the linear and angular jolt covectors are

F

F
i
= M v M
ij
v
j
= [

F
1
,

F
2
,

F
3
],

T

T
i
= I I
ij

j
= [

T
1
,

T
2
,

T
3
],
where
v = v
i
= [ v
1
, v
2
, v
3
]
t
, =
i
= [
1
,
2
,
3
]
t
,
are linear and angular jerk vectors.
In scalar form, the SE(3)jolt expands as
_
_
_

F
1
= p
1
m
2

3
v
2
+m
3
(
2
v
3
+v
3

2
) m
2
v
2

3
,

F
2
= p
2
+m
1

3
v
1
m
3

1
v
3
m
3
v
3

1
+m
1
v
1

3
,

F
3
= p
3
m
1

2
v
1
+m
2

1
v
2
v
2

1
m
1
v
1

2
,
_
_
_

T
1
=
1
(m
2
m
3
) (v
3
v
2
+v
2
v
3
) (I
2
I
3
) (
3

2
+
2

3
) ,

T
2
=
2
+ (m
1
m
3
) (v
3
v
1
+v
1
v
3
) + (I
1
I
3
) (
3

1
+
1

3
) ,

T
3
=
3
(m
1
m
2
) (v
2
v
1
+v
1
v
2
) (I
1
I
2
) (
2

1
+
1

2
) .
13
In this paragraph the overdots actually denote the absolute Bianchi (covariant) derivatives, so
that the jolts retain the proper covector character, which would be lost if ordinary time derivatives
are used. However, for simplicity, we stick to the same notation.
55
5.7 Symplectic Group in Hamiltonian Mechanics
Here we give a brief description of symplectic group (see [4, 8, 9]).
Let J =
_
0 I
I 0
_
, with I the n n identity matrix. Now, A L(R
2n
, R
2n
)
is called a symplectic matrix if A
T
J A = J. Let Sp(2n, R) be the set of 2n 2n
symplectic matrices. Taking determinants of the condition A
T
J A = J gives det A =
1, and so A GL(2n, R). Furthermore, if A, B Sp(2n, R), then (AB)
T
J(AB) =
B
T
A
T
JAB = J. Hence, AB Sp(2n, R), and if A
T
J A = J, then JA = (A
T
)
1
J =
(A
1
)
T
J, so J = (A1)
T
JA
1
, or A
1
Sp(2n, R). Thus, Sp(2n, R) is a group.
The symplectic Lie group
Sp(2n, R) =
_
A GL(2n, R) : A
T
J A = J
_
is a noncompact, connected Lie group of dimension 2n
2
+n. Its Lie algebra
sp(2n, R) =
_
A L(R
2n
, R
2n
) : A
T
J A = J = 0
_
,
called the symplectic Lie algebra, consists of the 2n 2n matrices A satisfying
A
T
J A = 0.
Consider a particle of mass mmoving in a potential V (q), where q
i
= (q
1
, q
2
, q
3
)
R
3
. Newtonian second law states that the particle moves along a curve q(t) in R
3
in
such a way that m q
i
= grad V (q
i
). Introduce the 3Dmomentum p
i
= m q
i
, and
the energy (Hamiltonian)
H(q, p) =
1
2m
3

i=1
p
2
i
+V (q).
Then
H
q
i
=
V
q
i
= m q
i
= p
i
, and
H
p
i
=
1
m
p
i
= q
i
, (i = 1, 2, 3),
and hence Newtonian law F = m q
i
is equivalent to Hamiltonian equations
q
i
=
H
p
i
, p
i
=
H
q
i
.
Now, writing z = (q
i
, p
i
),
J grad H(z) =
_
0 I
I 0
_
_
H
q
i
H
p
i
_
=
_
q
i
, p
i
_
= z,
56
so the complex Hamiltonian equations read
z = J grad H(z).
Now let f : R
3
R
3
R
3
R
3
and write w = f(z). If z(t) satises the complex
Hamiltonian equations then w(t) = f(z(t)) satises w = A
T
z, where A
T
= [w
i
/z
j
]
is the Jacobian matrix of f. By the chain rule,
w = A
T
J grad
z
H(z) = A
T
J Agrad
w
H(z(w)).
Thus, the equations for w(t) have the form of Hamiltonian equations with energy
K(w) = H(z(w)) i A
T
J A = J, that is, i A is symplectic. A nonlinear transfor-
mation f is canonical i its Jacobian matrix is symplectic. Sp(2n, R) is the linear
invariance group of classical mechanics.
6 Medical Applications: Prediction of Injuries
6.1 General Theory of MusculoSkeletal Injury Mechanics
The prediction and prevention of traumatic brain injury, spinal injury and musculo-
skeletal injury is a very important aspect of preventive medical science. Recently,
in a series of papers [28, 29, 30], we have proposed a new coupled loading-rate
hypothesis as a unique cause of all above injuries. This new hypothesis states that
the unique cause of brain, spinal and musculo-skeletal injuries is a Euclidean Jolt,
which is an impulsive loading that strikes any part of the human body (head, spine
or any bone/joint) in several coupled degrees-of-freedom simultaneously. It never
goes in a single direction only. Also, it is never a static force. It is always an
impulsive translational and/or rotational force coupled to some mass eccentricity.
This is, in a nutshell, our universal Jolt theory of all mechanical injuries.
To show this, based on the previously dened covariant force law, we have rstly
formulated the fully coupled NewtonEuler dynamics of:
1. Brains micro-motions within the cerebrospinal uid inside the cranial cavity;
2. Any local inter-vertebral motions along the spine; and
3. Any local joint motions in the human musculo-skeletal system.
Then, from it, we have dened the essential concept of Euclidean Jolt, which
is the main cause of all mechanical injuries. The Euclidean Jolt has two main
components:
1. Sudden motion, caused either by an accidental impact or slightly distorted
human movement; and
57
Figure 5: Human brain and its SE(3)group of microscopic three-dimensional mo-
tions within the cerebrospinal uid inside the cranial cavity.
2. Unnatural mass distribution of the human body (possibly with some added
masses), which causes some mass eccentricity from the natural physiological body
state.
What does this all mean? We will try to explain it in plain English. As we
live in a 3D space, one could think that motion of any part of the human body, ei-
ther caused by an accidental impact or by voluntary human movement, just obeys
classical mechanics in 6 degrees-of-freedom: three translations and three rotations.
However, these 6 degrees-of-freedom are not independent motions as it is suggested
by the standard term degrees-of-freedom. In reality, these six motions of any
body in space are coupled. Firstly, three rotations are coupled in the so-called ro-
tation group (or matrix, or quaternion). Secondly, three translations are coupled
with the rotation group to give the full Euclidean group of rigid body motions in
space. A simple way to see this is to observe someone throwing an object in the
air or hitting a tennis ball: how far and where it will y depends not only on the
standard projectile mechanics, but also on its local spin around all three axes
simultaneously. Every golf and tennis player knows this simple fact. Once the spin
is properly dened we have a fully coupled NewtonEuler dynamics to start with.
58
Figure 6: Human body representation in terms of SE(3)/SE(2)-groups of rigid-body
motion, with the vertebral column represented as a chain of 26 exibly-coupled
SE(3)-groups.
The covariant force law for any biodynamical system (which we introduced earlier
in our biodynamics books and papers, see our references in the cited papers above)
goes one step beyond the NewtonEuler dynamics. It states:
Euclidean Force covector eld =
Body mass distribution Euclidean Acceleration vector eld
This is a nontrivial biomechanical generalization of the fundamental Newtons
denition of the force acting on a single particle. Unlike classical engineering me-
chanics of multi-body systems, this fundamental law of biomechanics proposes that
forces acting on a multi-body system and causing its motions are fundamentally dif-
ferent physical quantities from the resulting accelerations. In simple words, forces
are massive quantities while accelerations are massless quantities. More precisely,
the acceleration vector eld includes all linear and angular accelerations of individual
body segments. When we couple them all with the total bodys mass-distribution
matrix of all body segments (including all masses and inertia moments), we get the
force co-vector eld, comprising all the forces and torques acting on the individual
body segments. In this way, we have dened the 6-dimensional Euclidean force for
an arbitrary biomechanical system.
Now, for prediction of injuries, we need to take the rate-of-change (or derivative,
59
Figure 7: Schematic latero-frontal view of the left knee joint. Although designed
to perform mainly exion/extension (strictly in the sagittal plane) with some re-
stricted medial/lateral rotation in the semi-exed position, it is clear that the knee
joint really has at least six-degrees-of-freedom, including three micro-translations.
The injury actually occurs when some of these microscopic translations become
macroscopic, which normally happens only after an external jolt.
with respect to time) of the Euclidean biomechanical force dened above. In this way,
we get the Euclidean Jolt, which is the sudden change (in time) of the 6-dimensional
Euclidean force:
Euclidean Jolt covector eld =
Body mass distribution Euclidean Jerk vector eld
And again, it consists of two components: (i) massless linear and angular jerks
(of all included body segments), and (ii) their mass distribution. For the sake of sim-
plicity, we can say that the mass distribution matrix includes all involved segmental
masses and inertia moments, as well as eccentricities or pathological leverages
from the normal physiological state.
Therefore, the unique cause of all brain, spine and musculo-skeletal injuries has
two components:
1. Coupled linear and angular jerks; and
2. Mass distribution with eccentricities.
60
In other words, there are no injuries in static conditions without any mass
eccentricities; all injuries are caused by mutually coupled linear and an-
gular jerks, which are also coupled with the involved human mass distri-
bution.
The Euclidean Jolt causes two forms of discontinuous brain, spine or musculo-
skeletal injury:
1. Mild rotational disclinations; and
2. Severe translational dislocations (or, fractures).
In the cited papers above, we have developed the soft-body dynamics of biome-
chanical disclinations and dislocations, caused by the Euclidean Jolt, using the
Cosserat multipolar viscoelastic continuum model.
Implications of the new universal theory are various, as follows.
A. The research in traumatic brain injury (TBI, see Figure 5) has so far iden-
tied the rotation of the brain-stem as the main cause of the TBI due to various
crashes/impacts. The contribution of our universal Jolt theory to the TBI research
is the following:
1. Rigorously dened this brain rotation as a mechanical disclination of the
brain-stem tissue modelled by the Cosserat multipolar soft-body model;
2. Showing that brain rotation is never uni-axial but always three-axial;
3. Showing that brain rotation is always coupled with translational disloca-
tions. This is a straightforward consequence of our universal Jolt theory.
These apparently obvious facts are actually radically new: we cannot separately
analyze rapid brains rotations from translations, because they are in reality always
coupled.
One practical application of the brain Jolt theory is in design of helmets. Briey,
a hard helmet saves the skull but not the brain; alternatively, a soft helmet pro-
tects the brain from the collision jolt but does not protect the skull. A good helmet
is both hard and soft. A proper helmet would need to have both a hard external
shell (to protect the skull) and a soft internal part (that will dissipate the energy
from the collision jolt by its own destruction, in the same way as a car saves its
passengers from the collision jolt by its own destruction).
Similarly, in designing safer car air-bags, the two critical points will be (i) their
61
placement within the car, and (ii) their soft-hard characteristics, similar to the
helmet characteristics described above.
B. In case of spinal injury (see Figure 6), the contribution of our universal Jolt
theory is the following:
1. The spinal injury is always localized at the certain vertebral or inter-vertebral
point;
2. In case of severe translational injuries (vertebral fractures or discus herniae)
they can be identied using X-ray or other medical imaging scans; in case of micro-
scopic rotational injuries (causing the back-pain syndrome) they cannot be identied
using current medical imaging scans;
3. There is no spinal injury without one of the following two causes:
a. Impulsive rotational + translational loading caused by either fast human
movements or various crashes/impacts; and/or
b. Static eccentricity from the normal physiological spinal form, caused by
external loading;
c. Any spinal injury is caused by a combination of the two points above:
impulsive rotational + translational loading and static eccentricity.
This is a straightforward consequence of our universal Jolt theory. We cannot
separately analyze translational and rotational spinal injuries. Also, there are no
static injuries without eccentricity. Indian women have for centuries carried bulky
loads on their heads without any spinal injuries; they just prevented any load ec-
centricities and any jerks in their motion.
The currently used Principal loading hypothesis that describes spinal injuries
in terms of spinal tension, compression, bending, and shear, covers only a small
subset of all spinal injuries covered by our universal Jolt theory. To prevent spinal
injuries we need to develop spinal jolt awareness: ability to control all possible im-
pulsive spinal loadings as well as static eccentricities.
C. In case of general musculo-skeletal injury (see Figure 7 for the particular case
of knee injury), the contribution of our universal Jolt theory is the following:
1. The injury is always localized at the certain joint or bone and caused by an
impulsive loading, which hits this particular joint/bone in several coupled degrees-
of-freedom simultaneously;
2. Injury happens when most of the body mass is hanging on that joint; for
example, in case of a knee injury, when most of the body mass is on one leg with
62
a semi-exed knee and then, caused by some external shock, the knee suddenly
jerks (this can happen in running, skiing, and ball games, as well as various
crashes/impacts); or, in case of shoulder injury, when most of the body mass is
hanging on one arm and then it suddenly jerks.
To prevent these injuries we need to develop musculo-skeletal jolt awareness. For
example, never overload a exed knee and avoid any kind of uncontrolled motions
(like slipping) or collisions with external objects.
6.2 Analytical Mechanics of Traumatic Brain Injury (TBI)
6.2.1 The SE(3)jolt: the cause of TBI
In this subsection we give a brief on TBI mechanics. For more details and references,
see [28].
In the language of modern dynamics, the microscopic motion of human brain
within the skull is governed by the Euclidean SE(3)group of 3D motions (see next
subsection). Within brains SE(3)group we have both SE(3)kinematics (consisting
of SE(3)velocity and its two time derivatives: SE(3)acceleration and SE(3)jerk)
and SE(3)dynamics (consisting of SE(3)momentum and its two time derivatives:
SE(3)force and SE(3)jolt), which is brains kinematics brains massinertia
distribution.
Informally, the external SE(3)jolt
14
is a sharp and sudden change in the SE(3)
force acting on brains massinertia distribution (given by brains mass and inertia
matrices). That is, a deltachange in a 3D forcevector coupled to a 3D torque
vector, striking the headshell with the brain immersed into the cerebrospinal uid.
In other words, the SE(3)jolt is a sudden, sharp and discontinues shock in all
6 coupled dimensions of brains continuous micromotion within the cerebrospinal
uid (Figure 5), namely within the three Cartesian (x, y, z)translations and the
three corresponding Euler angles around the Cartesian axes: roll, pitch and yaw.
If the SE(3)jolt produces a mild shock to the brain (e.g., strong head shake), it
causes mild TBI, with temporary disabled associated sensory-motor and/or cognitive
functions and aecting respiration and movement. If the SE(3)jolt produces a hard
14
The mechanical SE(3)jolt concept is based on the mathematical concept of higherorder tan-
gency (rigorously dened in terms of jet bundles of the heads conguration manifold), as follows:
When something hits the human head, or the head hits some external body, we have a collision.
This is naturally described by the SE(3)momentum, which is a nonlinear coupling of 3 linear New-
tonian momenta with 3 angular Eulerian momenta. The tangent to the SE(3)momentum, dened
by the (absolute) time derivative, is the SE(3)force. The second-order tangency is given by the
SE(3)jolt, which is the tangent to the SE(3)force, also dened by the time derivative.
63
shock (hitting the head with external mass), it causes severe TBI, with the total
loss of gesture, speech and movement.
The SE(3)jolt is the absolute timederivative of the covariant force 1form (or,
co-vector eld). The fundamental law of biomechanics is the covariant force law:
Force co-vector eld = Mass distribution Acceleration vectoreld,
which is formally written (using the Einstein summation convention, with indices
labelling the three Cartesian translations and the three corresponding Euler angles):
F

= m

, (, = 1, ..., 6)
where F

denotes the 6 covariant components of the external pushing SE(3)force


co-vector eld, m

represents the 66 covariant components of brains inertia


metric tensor, while a

corresponds to the 6 contravariant components of brains


internal SE(3)acceleration vector-eld.
Now, the covariant (absolute, Bianchi) time-derivative
D
dt
() of the covariant
SE(3)force F

denes the corresponding external striking SE(3)jolt co-vector


eld:
D
dt
(F

) = m

D
dt
(a

) = m

_
a

_
, (6)
where
D
dt
(a

) denotes the 6 contravariant components of brains internal SE(3)jerk


vector-eld and overdot () denotes the time derivative.

are the Christoels


symbols of the LeviCivita connection for the SE(3)group, which are zero in case
of pure Cartesian translations and nonzero in case of rotations as well as in the
fullcoupling of translations and rotations.
In the following, we elaborate on the SE(3)jolt concept (using vector and tensor
methods) and its biophysical TBI consequences in the form of brains dislocations
and disclinations.
6.2.2 SE(3)group of brains micromotions within the CSF
The brain and the CSF together exhibit periodic microscopic translational and ro-
tational motion in a pulsatile fashion to and from the cranial cavity, in the fre-
quency range of normal heart rate (with associated periodic squeezing of brains
ventricles). This micromotion is mathematically dened by the Euclidean (gauge)
SE(3)group.
In other words, the gauge SE(3)group of Euclidean micro-motions of the brain
immersed in the cerebrospinal uid within the cranial cavity, contains matrices of
the form
_
R b
0 1
_
, where b is brains 3D micro-translation vector and R is brains
3D rotation matrix, given by the product R = R

of brains three Eulerian


64
micro-rotations, roll = R

, pitch = R

, yaw = R

, performed respectively about


the xaxis by an angle , about the yaxis by an angle , and about the zaxis
by an angle ,
R

=
_
_
1 0 0
0 cos sin
0 sin cos
_
_
, R

=
_
_
cos 0 sin
0 1 0
sin 0 cos
_
_
, R

=
_
_
cos sin 0
sin cos 0
0 0 1
_
_
.
Therefore, brains natural SE(3)dynamics within the cerebrospinal uid is
given by the coupling of Newtonian (translational) and Eulerian (rotational) equa-
tions of micro-motion.
6.2.3 Brains natural SE(3)dynamics
To support our coupled loadingrate hypothesis, we formulate the coupled Newton
Euler dynamics of brains micro-motions within the sculls SE(3)group of motions.
The forced NewtonEuler equations read in vector (boldface) form
Newton : p M v = F +p , (7)
Euler : I = T+ +p v,
where denotes the vector cross product,
15
M M
ij
= diagm
1
, m
2
, m
3
and I I
ij
= diagI
1
, I
2
, I
3
, (i, j = 1, 2, 3)
are brains (diagonal) mass and inertia matrices,
16
dening brains massinertia
distribution, with principal inertia moments given in Cartesian coordinates (x, y, z)
by volume integrals
I
1
=
___
(z
2
+y
2
)dxdydz, I
2
=
___
(x
2
+z
2
)dxdydz, I
3
=
___
(x
2
+y
2
)dxdydz,
15
Recall that the cross product u v of two vectors u and v equals u v = uv sin n, where
is the angle between u and v, while n is a unit vector perpendicular to the plane of u and v such
that u and v form a right-handed system.
16
In reality, mass and inertia matrices (M, I) are not diagonal but rather full 3 3 positive
denite symmetric matrices with coupled mass and inertiaproducts. Even more realistic, fully
coupled massinertial properties of a brain immersed in (incompressible, irrotational and inviscid)
cerebrospinal uid are dened by the single non-diagonal 6 6 positivedenite symmetric mass
inertia matrix /
SE(3)
, the so-called material metric tensor of the SE(3)group, which has all
nonzero massinertia coupling products. In other words, the 6 6 matrix /
SE(3)
contains: (i)
brains own mass plus the added mass matrix associated with the uid, (ii) brains own inertia plus
the added inertia matrix associated with the potential ow of the uid, and (iii) all the coupling
terms between linear and angular momenta. However, for simplicity, in this paper we shall consider
only the simple case of two separate diagonal 3 3 matrices (M, I).
65
dependent on brains density = (x, y, z),
v v
i
= [v
1
, v
2
, v
3
]
t
and
i
= [
1
,
2
,
3
]
t
(where [ ]
t
denotes the vector transpose) are brains linear and angular velocity
vectors
17
(that is, column vectors),
F F
i
= [F
1
, F
2
, F
3
] and T T
i
= [T
1
, T
2
, T
3
]
are gravitational and other external force and torque co-vectors (that is, row vectors)
acting on the brain within the scull,
p p
i
Mv = [p
1
, p
2
, p
3
] = [m
1
v
1
, m
2
v
2
, m
2
v
2
] and

i
I = [
1
,
2
,
3
] = [I
1

1
, I
2

2
, I
3

3
]
are brains linear and angular momentum co-vectors.
In tensor form, the forced NewtonEuler equations (7) read
p
i
M
ij
v
j
= F
i
+
j
ik
p
j

k
, (i, j, k = 1, 2, 3)

i
I
ij

j
= T
i
+
j
ik

k
+
j
ik
p
j
v
k
,
where the permutation symbol
j
ik
is dened as

j
ik
=
_

_
+1 if (i, j, k) is (1, 2, 3), (3, 1, 2) or (2, 3, 1),
1 if (i, j, k) is (3, 2, 1), (1, 3, 2) or (2, 1, 3),
0 otherwise: i = j or j = k or k = i.
In scalar form, the forced NewtonEuler equations (7) expand as
Newton :
_
_
_
p
1
= F
1
m
3
v
3

2
+m
2
v
2

3
p
2
= F
2
+m
3
v
3

1
m
1
v
1

3
p
3
= F
3
m
2
v
2

1
+m
1
v
1

2
, (8)
Euler :
_
_
_

1
= T
1
+ (m
2
m
3
)v
2
v
3
+ (I
2
I
3
)
2

3

2
= T
2
+ (m
3
m
1
)v
1
v
3
+ (I
3
I
1
)
1

3

3
= T
3
+ (m
1
m
2
)v
1
v
2
+ (I
1
I
2
)
1

2
,
showing brains individual mass and inertia couplings.
17
In reality, is a 3 3 attitude matrix. However, for simplicity, we will stick to the (mostly)
symmetrical translationrotation vector form.
66
Equations (7)(8) can be derived from the translational + rotational kinetic
energy of the brain
18
E
k
=
1
2
v
t
Mv +
1
2

t
I, (9)
or, in tensor form
E =
1
2
M
ij
v
i
v
j
+
1
2
I
ij

j
.
For this we use the KirchhoLagrangian equations:
d
dt

v
E
k
=
v
E
k
+F, (10)
d
dt

E
k
=

E
k
+
v
E
k
v +T,
where
v
E
k
=
E
k
v
,

E
k
=
E
k

; in tensor form these equations read


d
dt

v
i E =
j
ik
(
v
j E)
k
+F
i
,
d
dt

i E =
j
ik
(

j E)
k
+
j
ik
(
v
j E) v
k
+T
i
.
Using (9)(10), brains linear and angular momentum co-vectors are dened as
p =
v
E
k
, =

E
k
,
or, in tensor form
p
i
=
v
i E,
i
=

i E,
with their corresponding time derivatives, in vector form
p =
d
dt
p =
d
dt

v
E, =
d
dt
=
d
dt

E,
or, in tensor form
p
i
=
d
dt
p
i
=
d
dt

v
i E,
i
=
d
dt

i
=
d
dt

i E,
18
In a fullycoupled NewtonEuler brain dynamics, instead of equation (9) we would have brains
kinetic energy dened by the inner product:
E
k
=
1
2
__
p

/
SE(3)
_
p

__
.
67
or, in scalar form
p = [ p
1
, p
2
, p
3
] = [m
1
v
1
, m
2
v
2
, m
3
v
3
], = [
1
,
2
,
3
] = [I
1

1
, I
2

2
, I
3

3
].
While brains healthy SE(3)dynamics within the cerebrospinal uid is given
by the coupled NewtonEuler microdynamics, the TBI is actually caused by the
sharp and discontinuous change in this natural SE(3) micro-dynamics, in the form
of the SE(3)jolt, causing brains discontinuous deformations.
6.2.4 Brains traumatic dynamics: the SE(3)jolt
The SE(3)jolt, the actual cause of the TBI (in the form of the brains plastic
deformations), is dened as a coupled Newton+Euler jolt; in (co)vector form the
SE(3)jolt reads
19
SE(3) jolt :
_
Newton jolt :

F = p p p ,
Euler jolt :

T = p v p v,
where the linear and angular jolt co-vectors are

F M v = [

F
1
,

F
2
,

F
3
],

T I = [

T
1
,

T
2
,

T
3
],
where
v = [ v
1
, v
2
, v
3
]
t
, = [
1
,
2
,
3
]
t
,
are linear and angular jerk vectors.
In tensor form, the SE(3)jolt reads
20

F
i
= p
i

j
ik
p
j

j
ik
p
j

k
, (i, j, k = 1, 2, 3)

T
i
=
i

j
ik

j

j
ik

j

k

j
ik
p
j
v
k

j
ik
p
j
v
k
,
in which the linear and angular jolt covectors are dened as

F

F
i
= M v M
ij
v
j
= [

F
1
,

F
2
,

F
3
],

T

T
i
= I I
ij

j
= [

T
1
,

T
2
,

T
3
],
where v = v
i
, and =
i
are linear and angular jerk vectors.
19
Note that the derivative of the crossproduct of two vectors follows the standard calculus
productrule:
d
dt
(u v) = u v +u v.
20
In this paragraph the overdots actually denote the absolute Bianchi (covariant) time-derivative
(6), so that the jolts retain the proper covector character, which would be lost if ordinary time
derivatives are used. However, for the sake of simplicity and wider readability, we stick to the same
overdot notation.
68
In scalar form, the SE(3)jolt expands as
Newton jolt :
_
_
_

F
1
= p
1
m
2

3
v
2
+m
3
(
2
v
3
+v
3

2
) m
2
v
2

3
,

F
2
= p
2
+m
1

3
v
1
m
3

1
v
3
m
3
v
3

1
+m
1
v
1

3
,

F
3
= p
3
m
1

2
v
1
+m
2

1
v
2
v
2

1
m
1
v
1

2
,
Euler jolt :
_
_
_

T
1
=
1
(m
2
m
3
) (v
3
v
2
+v
2
v
3
) (I
2
I
3
) (
3

2
+
2

3
) ,

T
2
=
2
+ (m
1
m
3
) (v
3
v
1
+v
1
v
3
) + (I
1
I
3
) (
3

1
+
1

3
) ,

T
3
=
3
(m
1
m
2
) (v
2
v
1
+v
1
v
2
) (I
1
I
2
) (
2

1
+
1

2
) .
We remark here that the linear and angular momenta (p, ), forces (F, T) and
jolts (

F,

T) are co-vectors (row vectors), while the linear and angular velocities
(v, ), accelerations ( v, ) and jerks ( v, ) are vectors (column vectors). This
bio-physically means that the jerk vector should not be confused with the jolt co-
vector. For example, the jerk means shaking the heads own massinertia matrices
(mainly in the atlantooccipital and atlantoaxial joints), while the joltmeans ac-
tually hitting the head with some external massinertia matrices included in the
hitting SE(3)jolt, or hitting some external static/massive body with the head
(e.g., the ground gravitational eect, or the wall inertial eect). Consequently,
the mass-less jerk vector represents a (translational+rotational) non-collision eect
that can cause only weaker brain injuries, while the inertial jolt co-vector represents
a (translational+rotational) collision eect that can cause hard brain injuries.
6.2.5 Brains dislocations and disclinations caused by the SE(3)jolt
Recall from introduction that for mild TBI, the best injury predictor is considered
to be the product of brains strain and strain rate, which is the standard isotropic
viscoelastic continuum concept. To improve this standard concept, in this subsec-
tion, we consider human brain as a 3D anisotropic multipolar Cosserat viscoelastic
continuum, exhibiting coupledstressstrain elastic properties. This non-standard
continuum model is suitable for analyzing plastic (irreversible) deformations and
fracture mechanics in multi-layered materials with microstructure (in which slips
and bending of layers introduces additional degrees of freedom, non-existent in the
standard continuum models.
The SE(3)jolt (

F,

T) causes two types of brains rapid discontinuous deforma-
tions:
1. The Newton jolt

F can cause micro-translational dislocations, or discontinuities
in the Cosserat translations;
2. The Euler jolt

T can cause micro-rotational disclinations, or discontinuities in
the Cosserat rotations.
69
To precisely dene brains dislocations and disclinations, caused by the SE(3)jolt
(

F,

T), we rst dene the coordinate co-frame, i.e., the set of basis 1forms dx
i
,
given in local coordinates x
i
= (x
1
, x
2
, x
3
) = (x, y, z), attached to brains center-
of-mass. Then, in the coordinate co-frame dx
i
we introduce the following set of
brains plasticdeformationrelated SE(3)based dierential pforms
21
:
the dislocation current 1form, J = J
i
dx
i
;
the dislocation density 2form, =
1
2

ij
dx
i
dx
j
;
the disclination current 2form, S =
1
2
S
ij
dx
i
dx
j
; and
the disclination density 3form, Q =
1
3!
Q
ijk
dx
i
dx
j
dx
k
,
where denotes the exterior wedgeproduct. These four SE(3)based dieren-
tial forms satisfy the following set of continuity equations:
= dJ S, (11)

Q = dS, (12)
d = Q, (13)
dQ = 0, (14)
where d denotes the exterior derivative.
In components, the simplest, fourth equation (14), representing the Bianchi iden-
21
Dierential pforms are totally skew-symmetric covariant tensors, dened using the exterior
wedgeproduct and exterior derivative. The proper denition of exterior derivative d for a pform
on a smooth manifold M, includes the Poincare lemma: d(d) = 0, and validates the general
Stokes formula
_
M
=
_
M
d,
where M is a pdimensional manifold with a boundary and M is its (p1)dimensional boundary,
while the integrals have appropriate dimensions.
A pform is called closed if its exterior derivative is equal to zero,
d = 0.
From this condition one can see that the closed form (the kernel of the exterior derivative operator
d) is conserved quantity. Therefore, closed pforms possess certain invariant properties, physically
corresponding to the conservation laws.
A pform that is an exterior derivative of some (p 1)form ,
= d,
is called exact (the image of the exterior derivative operator d). By Poincare lemma, exact forms
prove to be closed automatically,
d = d(d) = 0.
This lemma is the foundation of the de Rham cohomology theory
70
tity, can be rewritten as
dQ =
l
Q
[ijk]
dx
l
dx
i
dx
j
dx
k
= 0,
where
i
/x
i
, while
[ij...]
denotes the skew-symmetric part of
ij...
.
Similarly, the third equation (13) in components reads
1
3!
Q
ijk
dx
i
dx
j
dx
k
=
k

[ij]
dx
k
dx
i
dx
j
, or
Q
ijk
= 6
k

[ij]
.
The second equation (12) in components reads
1
3!

Q
ijk
dx
i
dx
j
dx
k
=
k
S
[ij]
dx
k
dx
i
dx
j
, or

Q
ijk
= 6
k
S
[ij]
.
Finally, the rst equation (11) in components reads
1
2

ij
dx
i
dx
j
= (
j
J
i

1
2
S
ij
) dx
i
dx
j
, or

ij
= 2
j
J
i
S
ij
.
In words, we have:
The 2form equation (11) denes the time derivative =
1
2

ij
dx
i
dx
j
of the
dislocation density as the (negative) sum of the disclination current S and
the curl of the dislocation current J.
The 3form equation (12) states that the time derivative

Q =
1
3!

Q
ijk
dx
i
dx
j

dx
k
of the disclination density Q is the (negative) divergence of the disclination
current S.
The 3form equation (13) denes the disclination density Q as the divergence
of the dislocation density , that is, Q is the exact 3form.
The Bianchi identity (14) follows from equation (13) by Poincare lemma and
states that the disclination density Q is conserved quantity, that is, Q is the
closed 3form. Also, every 4form in 3D space is zero.
From these equations, we can derive two important conclusions:
71
1. Being the derivatives of the dislocations, brains disclinations are higherorder
tensors, and thus more complex quantities, which means that they present a
higher risk for the severe TBI than dislocations a fact which is supported
by the literature (see review of existing TBImodels given in Introduction of
[28]).
2. Brains dislocations and disclinations are mutually coupled by the underlaying
SE(3)group, which means that we cannot separately analyze translational
and rotational TBIs a fact which is not supported by the literature.
For more medical details and references, see [28].
References
[1] Arnold, V.I., Mathematical Methods of Classical Mechanics. Springer, New
York, (1978)
[2] Abraham, R., Marsden, J., Foundations of Mechanics. Benjamin, Reading, MA,
(1978)
[3] Abraham, R., Marsden, J., Ratiu, T., Manifolds, Tensor Analysis and Applica-
tions. Springer, New York, (1988)
[4] Marsden, J.E., Ratiu, T.S., Introduction to Mechanics and Symmetry, A Basic
Exposition of Classical Mechanical Systems. (2nd ed), Springer, New York,
(1999)
[5] Ivancevic, V. Symplectic Rotational Geometry in Human Biomechanics. SIAM
Rev. 46(3), 455474, (2004)
[6] Ivancevic, V., Ivancevic, T., Human-Like Biomechanics, A Unied Mathe-
matical Approach to Human Biomechanics and Humanoid Robotics. Springer,
Berlin, (2005)
[7] Ivancevic, V., Ivancevic, T., Natural Biodynamics. World Scientic, Singapore
(2006)
[8] Ivancevic, V., Ivancevic, T., Geometrical Dynamics of Complex Systems, A
Unied Modelling Approach to Physics, Control, Biomechanics, Neurodynamics
and Psycho-Socio-Economical Dynamics. Springer, Dordrecht, (2006)
[9] Ivancevic, V., Ivancevic, T., Applied Dierential Geometry, A Modern Intro-
duction. World Scientic, Singapore, (2007)
72
[10] Switzer, R.K., Algebraic Topology Homology and Homotopy. (in Classics in
Mathematics), Springer, New York, (1975)
[11] de Rham, G., Dierentiable Manifolds. Springer, Berlin, (1984)
[12] Lang, S., Fundamentals of Dierential Geometry. Graduate Texts in Mathe-
matics, Springer, New York, (1999)
[13] Lang, S., Introduction to Dierentiable Manifolds (2nd ed.). Graduate Texts in
Mathematics, Springer, New York, (2002)
[14] Dieudonne, J.A., Foundations of Modern Analysis (in four volumes). Academic
Press, New York, (1969)
[15] Dieudonne, J.A., A History of Algebraic and Dierential Topology 1900-1960.
Birkh

auser, Basel, (1988)


[16] Spivak, M., Calculus on Manifolds, A Modern Approach to Classical Theorems
of Advanced Calculus. HarperCollins Publishers, (1965)
[17] Spivak, M., A comprehensive introduction to dierential geometry, Vol.I-V,
Publish or Perish Inc., Berkeley, (1970-75)
[18] Choquet-Bruhat, Y., DeWitt-Morete, C., Analysis, Manifolds and Physics (2nd
ed). North-Holland, Amsterdam, (1982)
[19] Choquet-Bruhat, Y., DeWitt-Morete, C., Analysis, Manifolds and Physics, Part
II, 92 Applications (rev. ed). North-Holland, Amsterdam, (2000)
[20] Bott, R., Tu, L.W., Dierential Forms in Algebraic Topology. Graduate Texts
in Mathematics, Springer, New York, (1982)
[21] Chevalley, C., Theory of Lie groups, Princeton Univ. Press, Princeton, (1946)
[22] Helgason, S., Dierential Geometry, Lie Groups and Symmetric Spaces. (2nd
ed.) American Mathematical Society, Providence, RI, (2001)
[23] Gilmore, R., Lie Groups, Lie Algebras and Some of their Applications (2nd
ed.), Dover, (2002)
[24] Fulton, W., Harris, J., Representation theory. A rst course, Graduate Texts
in Mathematics, Springer, New York, (1991)
[25] Bourbaki, N., Elements of Mathematics, Lie Groups and Lie Algebras, Springer,
(2002)
73
[26] Conway, J.H., Curtis, R.T., Norton, S.P., Parker, R.A., Wilson, R.A., Atlas
of Finite Groups: Maximal Subgroups and Ordinary Characters for Simple
Groups. Clarendon Press, Oxford, (1985)
[27] Schafer, R.D., An Introduction to Nonassociative Algebras. Dover, New York,
(1996)
[28] V.G. Ivancevic, New mechanics of traumatic brain injury, Cogn. Neurodyn.
3:281-293, (2009)
http://www.springerlink.com/content/p27023577564202h/?p=4351a9d0d76a4fd4b45d6720dad056f3&pi=8
[29] V.G. Ivancevic, New mechanics of spinal injury, IJAM, 1(2): 387401, (2009)
http://www.worldscinet.com/ijam/01/0102/S1758825109000174.html
[30] V.G. Ivancevic, New mechanics of generic musculo-skeletal injury, BRL,
4(3):273287, (2009)
http://www.worldscinet.com/brl/04/0403/S1793048009001022.html
74

You might also like